123
NOVEL APPROACHES TO AUTOMATED QUALITY CONTROL ANALYSES OF EDIBLE OILS BY FOURIER TRANSFORM INFRARED SPECTROSCOPY: DETERMINATION OF FREE FATTY ACID AND MOISTURE CONTENT Ahmed Ali AI-Alawi September 2005 Department of Food Science and Agricultural Chemistry McGill University, Montreal A thesis submitted to the Faculty of Graduate Studies and Research in partial fulfillment of the requirements of the degree of Doctor of Philosophy ©Ahmed Ali AI-Alawi, 2005

Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

  • Upload
    others

  • View
    7

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

NOVEL APPROACHES TO AUTOMATED QUALITY CONTROL

ANALYSES OF EDIBLE OILS BY FOURIER TRANSFORM

INFRARED SPECTROSCOPY: DETERMINATION OF FREE

FATTY ACID AND MOISTURE CONTENT

Ahmed Ali AI-Alawi

September 2005

Department of Food Science and Agricultural Chemistry

Mc Gill University, Montreal

A thesis submitted to the Faculty of Graduate Studies and Research in partial fulfillment

of the requirements of the degree of Doctor of Philosophy

©Ahmed Ali AI-Alawi, 2005

Page 2: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

1+1 Library and Archives Canada

Bibliothèque et Archives Canada

Published Heritage Branch

Direction du Patrimoine de l'édition

395 Wellington Street Ottawa ON K1A ON4 Canada

395, rue Wellington Ottawa ON K1A ON4 Canada

NOTICE: The author has granted a non­exclusive license allowing Library and Archives Canada to reproduce, publish, archive, preserve, conserve, communicate to the public by telecommunication or on the Internet, loan, distribute and sell theses worldwide, for commercial or non­commercial purposes, in microform, paper, electronic and/or any other formats.

The author retains copyright ownership and moral rights in this thesis. Neither the thesis nor substantial extracts from it may be printed or otherwise reproduced without the author's permission.

ln compliance with the Canadian Privacy Act some supporting forms may have been removed from this thesis.

While these forms may be included in the document page cou nt, their removal does not represent any loss of content from the thesis.

• •• Canada

AVIS:

Your file Votre référence ISBN: 978-0-494-25088-4 Our file Notre référence ISBN: 978-0-494-25088-4

L'auteur a accordé une licence non exclusive permettant à la Bibliothèque et Archives Canada de reproduire, publier, archiver, sauvegarder, conserver, transmettre au public par télécommunication ou par l'Internet, prêter, distribuer et vendre des thèses partout dans le monde, à des fins commerciales ou autres, sur support microforme, papier, électronique et/ou autres formats.

L'auteur conserve la propriété du droit d'auteur et des droits moraux qui protège cette thèse. Ni la thèse ni des extraits substantiels de celle-ci ne doivent être imprimés ou autrement reproduits sans son autorisation.

Conformément à la loi canadienne sur la protection de la vie privée, quelques formulaires secondaires ont été enlevés de cette thèse.

Bien que ces formulaires aient inclus dans la pagination, il n'y aura aucun contenu manquant.

Page 3: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

Short title:

FTIR ANAL YSIS OF EDIBLE OILS FOR FREE FATTY ACIDS AND MOISTURE

Page 4: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

ABSTRACT

Three new quantitative Fourier transform infrared (FTIR) spectroscopie methods

were developed for the analysis of edible oils: two procedures to measure free fattyacids

(FFA) and one to measure moisture (HzO), the latter two methods ultimately being

automated and implemented on an auto-sampler equipped FTIR spectrometer. The

methods developed for FF A determination both convert FF As to their carboxylate salts

by me ans of acid/base reaction without causing oil saponification, one approach using 1-

propanol, an oil-miscible solvent, and the other using methanol, an oil-immiscible solvent

into which the FF A salts are extracted. The first method involves splitting oil samples

into two halves, with one half treated with propanol containing base and the other half

with propanol only. The spectra of each half is coHected and a differential spectrum

obtained, from which quantization is performed. The methanol procedure simply involves

extracting FF A into methanol containing a weak base and quantitating the FF A salts

produced. Both FF A methods determine the FF A content by measuring the v (COO")

absorbance at ~1570 cm-I relative to a reference wavelength of 1820 cm- I from a

differential spectrum relative to the solvent, the extraction procedure being superior in

terms ofboth speed and sensitivity, being able to measure FFA levels down to ~0.001%.

The method developed for moisture determination involves extracting water in edible oils

into dry acetonitrile and then quantitating it by measuring the absorbance of the OH

stretching band (3629 cm-I) and/or the HOH bending band (1631 cm- I

). AH three

methods were validated by standard addition experiments, evaluated for potential

interferences, and, in the case of FF A determination, compared to the performance of

AOCS official methods. The results indicated that the extraction-based procedures were

superior to conventional wet chemical methods in both sensitivity and reproducibility.

The FF A and H20 extraction procedures were subsequently automated by connecting an

auto-sampler to the FTIR spectrometer and developing procedures and software

algorithms to enable the analysis of up to 100 samples/h. The methods developed and

implemented are a substantive improvement over conventional methods for the analysis

of FF A and HzO in edible oils and provide a means by which QC and process

laboratories can analyze large volumes of edible oils for these two important parameters.

Page 5: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

RESUME

Trois nouvelles méthodes d'analyse quantitative utilisant la spectroscopie à

infrarouge à transformation de Fourier (FTIR) ont été développées afin d'analyser les

huiles comestibles; deux méthodes pour mesurer les acides gras libres (AGL) et une

méthode pour mesurer l'humidité (H20), ces méthodes ont été enfin automatisées et

mises en application sur un spectromètre FTIR équipé d'un auto échantillonneur.

Les deux méthodes développées pour la détermination des AGL transforment les

AGL en leur sels de carboxylate par le moyen de la réaction acide/base sans causer de

saponification à l 'huile, la première méthode utilise le propanol, un solvant miscible avec

l'huile alors que la deuxième méthode utilise le méthanol, un solvant immiscible avec

l'huile et dans lequel les sels des AGL sont extraits. La première méthode fait appel à la

division des échantillons d'huile en deux moitiés, une moitié est traitée par une base

dissous dans le propanol et l'autre moitié est traitée avec du propanol uniquement. Le

spectre de chaque moitié est enregistré et un spectre différentiel est obtenu permettant

ainsi la quantification. La méthode utilisant le méthanol fait simplement appel à

l'extraction des AGL dans une solution de méthanol contenant une base faible et à la

quantification des sels des AGL ainsi produits. Les deux méthodes déterminent le contenu en AGL en mesurant l'absorbance du v (COO) à 1570 cm- I respectif à une

longueur d'onde de référence de 1820 cm-1 à partir d'un spectre différentiel respectif au

solvant, elles sont capables de mesurer des niveaux des AGL aussi bas que ~0.001 %. Une

méthode pour la détermination de l'humidité a éte développée en extrayant l'eau dans les

huiles comestibles dans de l'acetonitrile sec et le quantifiant par la suite en mesurant

l'absorbance de la bande d'élongation OH (3629 cm-I) et/ou de la bande de déformation

HOH (1631 cm- I). Les trois méthodes furent ensuite validées grâce à l'utilisation de

techniques d'addition standard, evaluées pour les interférences potentielles et dans les cas

de la détermination des AGL comparées à la performance des méthodes officielles de

l'AOCS. Les résultats ont montré que les procédures basées sur l'extraction étaient

supérieures aux méthodes chimiques conventionnelles en termes de sensibilité et

reproductibilité. Les méthodes d'extraction des AGL et du H20 furent ensuite

automatisées en connectant un auto-echantillonneur au spectromètre FTIR et en

développant des procédures et des algorithmes pour logiciels qui facilitent l'analyse

jusqu'à 100 echantillons/ho Les méthodes développées et mises en application constituent

une amélioration considérable par rapport aux méthodes conventionnelles d'analyse des

AGL et H20 dans les huiles comestibles et fournissent un moyen par lequel le CQ et les

laboratoires des processus peuvent analyser de larges volumes d'huiles comestibles pour ces deux importants paramètres.

11

Page 6: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

ACKNOWLEDGMENTS

l would like to express my deepest gratitude to Professor F. R. van de Voort, my

supervisor, for the time, guidance and encouragement that he has given me throughout

my PhD program; his exceptional qualities as a research supervisor have helped make

this thesis possible. l wish to acknowledge my sincere gratitude to Dr. Jacqueline

Sedman for her collaboration, valued advice and input and to thank Drs. William

Marshall, Selim Kermasha, Varoujan Yaylayan and Ashraf Ismail for their guidance,

advice and use of analytical instrumentation.

Very special thanks to Sensei John Kalaidopoulos, my senior karate teacher,

who taught me in practice the true meaning of karate, for his support and understanding.

Very warm thanks go to Sempai Stephen, Sempai Tony, my best friend Denies and aIl

the karate students at the West Island Kyokushin Karate for their kindness, friendship

and continuous help. l also wish to extend my warm thanks to my colleagues at the Food

Science and Agricultural Chemistry Department and to my friends in Sainte Anne de

Bellevue.

Above aIl, my deepest gratitude goes to my parents, family and my beloved

woman for their continuous love, support and encouragement, with my personal thanks to

Sultan Qaboos University (SQU) and Embassy of Sultanate of Oman, Washington DC,

for the financial and academic support that they given me throughout my study in

Canada.

iii

Page 7: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

CONTRIBUTIONS OF AUTHORS

Chapters 3-6 of this thesis comprise of the text of published or submitted papers

listed below. The author of this thesis was responsible for the concept, design of

experiments, experimental work, and manuscript preparation. Dr. van de Voort is thesis

supervisor and had direct advisory input as the work as it progressed. Dr. Sedman

provided advice relative to spectral interpretation, chemometrics and editorial assistance

in the final stages of paper submission. Mr. Ghetler was responsible for the UMPlRE

software platform on which FF A and H20 algorithms were developed for the automated

methods discussed in Chapter 6.

List of the publications reported in the thesis:

AI-Alawi, A., van de Voort, F. R and Sedman, J. New FTIR Method for the

Determination of Free Fatty Acids in Oils. J. Am. Oil Chem. Soc., 81(5), 441-446 (2004).

AI-Alawi, A., van de Voort, F. R and Sedman, J. A New FTIR Method for the Analysis

ofLow Levels ofFFA in Refined Edible Oils. Spectrosc. Lett. 38(4-5) 389-403 (2005).

AI-Alawi, A., van de Voort, F. Rand Sedman, J. A New FTIR Method for the

Determination of Low Levels of Moisture in Edible Oils. Appl. Spectrosc., 59(10), 1295-

1299 (2005).

AI-Alawi, A., van de Voort, F. R, Sedman, J. and Ghetler, A., Automated FTIR Analysis

of Free Fatty Acids and Water in Edible Oils. Journal of the Association for Laboratory

Automation. (in press).

IV

Page 8: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

CONTRIBUTIONS TO KNOWLEDGE

1. Designed new analytical approaches for edible oil analysis to provide the

industry with practical and readily implemented methodology.

New analytical strategies were developed to compensate for matrix effects arising

in FTIR analysis of edible oils that are not readily modeled by chemometrics,

thereby enhancing analytical accuracy and sensitivity, to facilitate sample

handling, and to allow automation of the analysis.

2. Developed the first quantitative FTIR method for FF A analysis in edible oil

that avoids saponification and is independent of oH type.

3.

Quantification was successfully achieved through the use of carefully selected

weak bases which converted FF A to their salts without causing saponification of

the oil. Generalization of the FTIR method so that matrix effects were eliminated

and the spectral response was independent of oil type was achieved via

differential spectroscopy.

Developed a sensitive and simple method for determination of low levels of

FF A in refined edible oils.

The sensitivity of the FF A method was extended by treating oils with an oil­

immiscible solvent containing a weak base to convert the FF A to their carboxylate

salts and concentrate the salts in a small volume. This procedure greatly enhanced

the sensitivity of the FTIR FF A analysis as well as eliminating the requirement to

prepare two samples to obtain one analytical result as in the first method devised.

4. Developed a sensitive, simple and accurate method for the quantitative

determination of moisture content in edible oils.

An FTIR-based instrumental method for the determination of moi sture was

developed, modeled on the extraction procedure developed for low-FF A

determination, but without any requirement to carry out a stoichiometric reaction.

The method eliminates most of the analytical problems that are commonly

associated with moisture quantification in edible oils and provides a viable

alternative to the widely used Karl Fischer method.

v

Page 9: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

5. Automated the extraction-based FF A and moisture analysis methods to carry

out high-speed, high-volume analyses of these measures for QC and process

laboratory purposes.

The extraction-based FF A and moisture methods developed were refined and

implemented to operate on an FTIR integrated with an auto-sampler and micro­

pump to allow for automation of these methods. The system was developed,

programmed, optimized and validated and was shown to be able to analyze up to

100 samples/h.

VI

Page 10: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

TABLE OF CONTENTS

ABsTIti\<:T-------------------------------------------------------------------------------------------i

~SlJ~E----------------------------------------------------------------------------------------------ii

AC~O~LEI>(;~ENTS------------------------------------------------------------------------iii

<:ONTRIBlJTIONS OF A lJTH ORS ----------------------------------------------------------- iv

<:0 NTRIBlJTI ONS TO~ 0 ~LEI>(; E ------------------------------------------------------ v

TABLE OF CONTENTS------------------------------------------------------------------------ vii

LIST OF FI(;lJRES --------------------------------------------------------------------------------x

LIST OF TABLES-------------------------------------------------------------------------------- xii

LIST OF ABBREVIATI 0 NS ----------------------------------------------------------------- xiii CHAPTERI

(;ENEIti\L INTROI>lJCTION------------------------------------------------------------------- 1 1.1. CONTEXT OF THE RESEARCH ---------------------------------------------------- 1

1.2. RA TIONALE OF THE WORK -------------------------------------------------------- 3

1.3 . REFERENCES --------------------------------------------------------------------------- 4

<:HAPTER2

LITEIti\TlJ~ ~VIE~ -------------------------------------------------------------------------5 2.1. FREE F ATTY ACIDS (FF A) ANAL YSIS------------------------------------------- 5

2.1.1. Introduction ------------------------------------------------------------------------- 5

2.1.2. Historical Perspective-------------------------------------------------------------- 6

2.1.3. Colorimetric Methods-------------------------------------------------------------- 6

2.1.4. Electrochemical Methods ------------------------------------------------------- Il

2.1.5. Thermometric Methods ---------------------------------------------------------- 14

2.1.6. CUITent Status of FF A Analysis ------------------------------------------------ 14

2.1.7. Infrared Spectroscopic Methods ----------------------------------------------- 15

2.2. MOISTURE CONTENT ANAL YSIS ---------------------------------------------- 17

2.2.1. Introduction ----------------------------------------------------------------------- 17

2.2.2. Karl Fischer Method ------------------------------------------------------------- 19

2.2.3. Infrared Methods ----------------------------------------------------------------- 27

2.3. CONCLUSION ------------------------------------------------------------------------- 31

2.4. REFERENCES ------------------------------------------------------------------------- 31

vu

Page 11: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

CHAPTER3

NEW FTIR METHOD FOR THE DETERMINATION OF FREE FATTY AC ID IN

OI~~ ------------------------------------------------------------------------------------------------- 39

3 .1. AB~TRA CT ----------------------------------------------------------------------------- 39

3.2. ~TFt()l)lJCTI()~---------------------------------------------------------------------- 39

3.3. EXPEFtIME~T AL PFt()CEl)lJFtE -------------------------------------------------- 42

3.4. FtESlJLTS A~I) I)ISClJSSI()~ ----------------------------------------------------- 44

3.4.1. Analytical Concepts-------------------------------------------------------------- 44

3.4.2. Calibration and Stability of the Fteaction ------------------------------------- 45

3.4.3. Evaluation and Validation------------------------------------------------------- 48

3.5. A CKN () WLEI)GME~T -------------------------------------------------------------- 52

3. 6. FtEFEFtE~ CES ------------------------------------------------------------------------- 52

lJRID(;E---------~----------------------------------------------------------------------------------- 54

CHAPTER4

A NEW METHOD FOR THE ANA~Y~I~ OF ~OW ~EVE~~ OF FFA IN

REFINED EDIlJ~E OI~~----------------------------------------------------------------------- 55

4.1. ABSTRACT ----------------------------------------------------------------------------- 55

4.2. ~TFt()l)lJCTI()~---------------------------------------------------------------------- 55

4.3. METH()I)() L()G Y --------------------------------------------------------------------- 59

4.4. FtESlJLTS A~I) I)ISClJSSI()~ ----------------------------------------------------- 60

4.4.1. Calibration ------------------------------------------------------------------------- 63

4.4.2. V alidation -------------------------------------------------------------------------- 64

4.5. C() ~ CL lJSI () ~ ------------------------------------------------------------------------- 69

4.6. ACK~()WLEI)GME~T -------------------------------------------------------------- 69

4.7. FtEFEFtE~CES ------------------------------------------------------------------------- 69

lJRID(;E--------------------------------------------------------------------------------------------- 72

CHAPTER5

A NEW FTIR METHOD FOR THE DETERMINATION OF MOI~TURE IN

EDIlJ~E OI~~------------------------------------------------------------------------------------- 73

5.1. ABSTRACT ----------------------------------------------------------------------------- 73

5.2. ~TFt()l)lJCTI()~---------------------------------------------------------------------- 73

Vlll

Page 12: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

5.3. 1V1~lllI()I)()~()CJ): --------------------------------------------------------------------- 75

5.4. RESU~llS ANI) I)ISCUSSION ----------------------------------------------------- 78

5.4.1. CJeneral Considerations ---------------------------------------------------------- 78

5.4.2. Calibration I)evelopment -------------------------------------------------------- 80

5.4.3. Consideration of Interfering Constituents ------------------------------------ 82

5.4.4. Factors Affecting Analytical Accuracy --------------------------------------- 84

5.5. A CKN () W~~I)CJ~1V1~NllS----------------------------------------------------------- 85

5 .6. REF~RENC~S ------------------------------------------------------------------------- 85

1JFlII>(;IG--------------------------------------------------------------------------------------------- 87 CHAPTIGR6

AUTOMATIGI> FTIR ANALYSIS OF FRIGIG FATTY ACII>S OR MOISTURIG IN

IGI>I1JLIG ()ILS------------------------------------------------------------------------------------- 88

6.1. ABSllRACll ----------------------------------------------------------------------------- 88

6.2. INllFt()I)UCllI()N---------------------------------------------------------------------- 88

6.3. 1V1~lllI () I)() ~OCJ): --------------------------------------------------------------------- 88

6.4. RESU~ Ils ANI) I)ISCUSSI()N ----------------------------------------------------- 92

6.4.1. System ()verview and Software ------------------------------------------------ 93

6.4.2. System ~valuation and Validation --------------------------------------------- 99

6.5. C()NC~USI()N ------------------------------------------------------------------------ 101

6.6. A CKN () W~~I)CJ 1V1~Nll ------------------------------------------------------------- 1 02

6.7. REF~~NC~S ------------------------------------------------------------------------ 102

CHAPTIGR 7

(;IGNIGRAL CON CL USI ON ------------------------------------------------------------------- 104

IX

Page 13: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

LIST OF FIGURES

Figure 2.1. Schematic diagram of the copper soap method for FF A analysis introduced

by Baker in 1964 ...................•...............•.•........•..........•............•...•..........•...•.... 7

Figure 2.2. Flow diagram of the FIA instrument for determination of FF A in oi! samples

as proposed by Ekstrom and Rhee .....................................................•............. 8

Figure 2.3. Flow diagram of the FIA instrument for determination of FF A in oil samples

as proposed by Canham .•.....•.•...••.•.•.•.••.•.•.•.••••.•..•.•..••.••••••.••••••..••••.••..••••••.•...• 9

Figure 2.4. Flow diagram of the FIA instrument for determination of FF A in oil samples

as proposed by Linarers et al. in 1989 ........................................................... 11

Figure 2.5. Metrohm auto-titrator instrument equipped with autosampler for FF A

analysis .......................................................................................................... 13

Figure 2.6. Shift in FF A band after conversion to carboxylate salt. FF A shows band at

1711 cm-1 (A); the band shifts to 1570 cm-1 after addition of base (B) ..•...... 16

Figure 2.7. FTIR spectrum ofpure water .......................•..•.......•..•.••.••.•.••.•..•.•••.•.•.•.•.••.•.• 28

Figure 3.1. Schematic diagram illustrating the sample preparation procedure and the

steps in the analytical protocolleading to the differential spectrum ...•...•..•.• 43

Figure 3.2. Potassium phthalimide as a reagent to carry out the acid/base reaction ........ 46

Figure 3.3. Composite calibration curve obtained from the differential spectra of three

independent sets of hexanoic acid-spiked soybean oi! standards of the type

illustrated in Figure 3.1 ..............................................................•.•.....•...•..•.•. 47

Figure 3.4. Graphical comparison of results obtained by the AOCS and FTIR methods

for the oils that were subjected to standard addition of FF A mixture, relative

to the gravimetrical data ................................................................................. 50

Figure 4.1. Schematic diagram of the sample preparation and analytical protocol. ........ 61

Figure 4.2. DifferentiaI spectra for sodium hydrogen cyanamide in methanol.. .............. 62

Figure 4.3. DifferentiaI spectra for soybean oil spiked with palmitic acid (0.0 - 0.1 %)

after carrying out the acidlbase reaction ....•..•.........................•..•................... 64

Figure 4.4. Calibration curve for %FF A in oïl obtained from the differential spectra in

Figure 4.2. The %FFA is expressed as % oleic acid (w/w) ........................... 65

Figure 5.1. Schematic diagram of the FTIR sample handling system ............................. 76

x

Page 14: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

,~,

Figure 5.2. Schematic diagram of the sample preparation procedure for moisture analysis

of edible oils and the FTIR spectral analytical protocol. ............................... 77

Figure 5.3. Spectra of water/acetonitrile standards (0, 400, and 800 ppm added water)

(upper series) and the corresponding spectra obtained after subtraction of the

spectrum of the acetonitrile employed to prepare the standards (lower series) .

....................................................................................................................... 79

Figure 5.4. DifferentiaI spectra obtained after extraction of water from water/canola oil

standards (0-2000 ppm water) into acetonitrile, recording the spectrum of the

acetonitrile layer, and ratioing the spectrum against that of the acetonitrile

extraction sol vent ........................................................................................... 81

Figure 5.5. DifferentiaI spectra of acetonitrile spiked with hexanol (A), glycerol (B),

tert-butyl hydroperoxide (C), oleic acid (D), and water (shaded) .............•..• 83

Figure 6.1. General analytical protocol for FF A and H20 analysis of edible oils by FTIR

spectroscopy .................................................................................................. 92

Figure 6.2. The FTIRIauto-sampler system used for the analysis ofFFA or moisture .... 94

Figure 6.3. Auto-sampler needle punching through the Mylar liner. ............................... 94

Figure 6.4. The user interface of UMPIRE® Pro pro gram. A and B, the "General" and

"Scanning/Sampling" tabs for the "Modify Methods" submenu, respectively;

C, "Peak/Area Position" tab for the "Components" submenu ....................... 95

Figure 6.5. DifferentiaI spectra of canola oil standards spiked with water (0-2000 ppm)

after extraction with acetonitrile .................................................................... 98

Figure 6.6. DifferentiaI spectra of canola oil standards spiked with palmitic acid (0.0-

0.1 % w/w) after extraction with methanol containing NaHNCN .•..•........•.•.. 98

Figure 6.7. Plots for moisture content vs. amounts added for the validation run ........... 100

Figure 6.8. Plots for FF A vs. amounts added for the validation run .............................. 101

Xl

Page 15: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

LIST OF TABLES

Table 1.1. Data Entry Report for Cottonseed Oil (2001-2002) ........................................ 2

Table 3.1. Results of Triplicate Analyses of Oils Spiked with Various Amounts an FF A

Mixture by the AOCS Titrimetric and FTIR Methods .................................. 49

Table 3.2. Results of Triplicate Analysis of Smalley Check Samples by AOCS Method

and FTIR Method .......................................................................................... 51

Table 4.1. FTIR Analyses Based on the Use of Stoichiometric Reactions for "Signal

Transduction" ................................................................................................ 57 .

Table 4.2. Mean Difference (MD) and Standard Deviation of the Difference (SDD) for

Accuracy (a) for the Titrimetric and FTIR Procedures vs. Gravimetrie

Addition and for Reproducibility (r) for Duplicate Analyses Carried Out by

Titrimetric and FTIR Analyses, Respectively ............................................... 67

Table 4.3. Results of Triplicate Analyses of Locally Purchased Oil Samples by the

IUPAC Reference (Titrimetric) Method and the FTIR Method .................... 68

Table 6.1. FF A and Moisture Content of Various Edible Oils (mean of triplicate

analyses). . ...................................................................................................... 99

Xll

Page 16: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

LIST OF ABBREVIATIONS

I-PrOH n-Propanol

AN Acid number

AOAC Association of Official Analytical Chemists

AOCS American Oil Chemists' Society

ASTM American Society for Testing and Materials

ATR Attenuated total reflectance

BR Blank reagent

DMP Dimethoxypropane

DMSO Dimethyl sulfoxide

DTGS Deuterated triglycine sulfate

EDTA Ethylenediaminetetraacetic acid

FFA Free fatty acid

FIA Flow injection system

FTIR Fourier transform infrared spectroscopy ~--

IUPAC International Union of Pure and Applied Chemistry

K-Phthal Potassium phthalmide

LIMS Laboratory information management system

MD Mean difference

MeOH Methanol

MLR Multiple linear regression

NIR N ear infrared

NSERC Natural Sciences and Engineering Research Council

OD Optical density

PLS Partialleast-squares regression

ppm Parts per million

PV Peroxide value

QC Quality control

RBD Refined, bleached and deodorized oil

/'~--" ROOH Hydroperoxide

RR Reactive reagent

Xll1

Page 17: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

,/ -, ~ SD

SDD

SEFIA

TAG

TEk

TPP

TPPO

UMPIRE®

w/v

w/w

ZnSe

Z-reg

Standard deviation

Standard deviation of the differences

Solvent extraction flow injection system

Triacylglycerol

Triethanolamine

Triphenylphosphine

Triphenylphosphine oxide

Universal Method Platform for InfraRed Evaluation

weight per volume

weight per weight

Zinc selenide

Regression through the origin

XIV

Page 18: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

r·· CHAPTERI

GENERAL INTRODUCTION

1.1. CONTEXT OF THE RESEARCH

Fats and oils represent a major and essential portion of our diet, and the fats and

oils industry is one ofthe large st sectors of the food industry. According to Euromonitor

International, consumer expenditure for fats and oils in 2005 was US $ 122.6 billion (1),

comprising - 0.5% of the total consumer expenditures; this amount does not include by­

products such as soap and meal cake. Because fats and oils are valuable and economically

important food commodities, a variety of professional organizations have set standards

for the edible oil industry in relation to determining their quality, functionality and the

parameters affecting their economic value. Several prominent organizations vie to play

official roles in this regard, including the American Oil Chemists' Society (AOCS) (2),

the Association of Official Analytical Chemists (AOAC) and the International Union of

Pure and Applied Chemistry (IUP AC) (3). Methods and standards adopted by these

organizations are considered to be "official" methods, most being similar, albeit not

identical in their detail. A wide range of official analytical methods exist for the routine

evaluation and characterization of fats and oils, examples being iodine value (degree of

unsaturation), saponification value (average molecular weight), moisture content,

peroxide value, anisidine value and free fatty acid (FF A) content (4), to name just a few.

Most of the official procedures noted above are titrimetric methods developed in

the 1930-40's using color indicators and many are still in routine use today. AlI are

workable and are standardized; however, the lack of skilled labor and the

safety/environmental issues associated with toxic solvents and reagents make these

methods problematic in modem times. Apart from these drawbacks, most of these

methods are mediocre in terms of accuracy, reproducibility and analytical range relative

to newer instrumental methods. The po or accuracy is often associated with their non­

specificity, the reliance on accurate color perception by the analyst or problematic dark

samples, the latter obscuring the end point. Table 1.1 shows an example of an inter­

laboratory study of FF A and moi sture results obtained for the same sample of cottonseed

oil (5) as reported by 10 different certified laboratories using the same official methods.

1

Page 19: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

Table 1.1. Data Entry Report for Cotionseed Oil (2001-2002). Adopted from ref. (5)

Analyst FFAa Moistureb

1 0.14 0.10

2 0.15 0.05

3 0.16 0.09

4 0.15 0.10

5 0.07

6 0.04 0.09

7 0.17 0.04

8 0.05 0.08

9 0.18 0.06

a % w/w expressed as oleic acid; b % w/w

Such variations can increase even further as a result of biases which may

interpose themselves when "official" methods from different organizations are used (e.g.

AOCS vs. IUPAC), as each has different procedure. For environmental, sample

throughput and accuracy reasons, there has been an ongoing effort to develop new

instrumental approaches to titrimetric based analyses, focusing on automated titrators and

continuous flow injection systems using various potentiometric and colorimetric

detection systems.

An alternative to titrimetric analyses is the application of spectroscopic techniques

to determine oil quality parameters based on spectral characteristics of the sample. In

particular, a wide range of oil parameters can now be analyzed by FTIR spectroscopy (6),

providing alternative approaches to their assessment. The Mc Gill IR Group has been at

the forefront of the development of instrumental methods for the analysis of edible oils

based on Fourier transform infrared (FTIR) spectroscopy. Whereas infrared spectroscopy

.~... was generally regarded in the past as a qualitative technique, providing functional group

information, it has now evolved into a quantitative spectroscopie procedure, supported by

2

Page 20: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

powerful chemometrics. Thus quantitative work with FTIR instrumentation can now

approach the simplicity of using UV/visible spectroscopy, while providing much more

specific spectral information in relation to the molecular species targeted for analysis.

However, for complex samples such as edible oils there are certain inherent limitations to

FTIR analysis such as matrix effects, overlapping bands, hydrogen bonding effects, etc.

Chemometric techniques such as partial least squares regression can overcome sorne of

these issues, but not aIl of them. Method development by the McGill IR Group has

undergone an analytical evolution which exploits a range of strategies to obtain ever

better sensitivity, accuracy or speed. The ultimate goals are to improve sensitivity and

accuracy and ultimately to standardize and automate edible oil analysis by FTIR

spectroscopy.

1.2. RATIONALE OF THE WORK

Of the variety of methods one could tackle with the above goals in mind, FF A

analysis was my personal choice, based on the thousands of titrations which 1 had the

misfortune to carry out in an oil processing plant, being a good motivator. Although FTIR

methods for FF A determination have been previously developed, they suffer from several

shortcomings. Similarly, there are weIl known problems associated with the titrimetric

determination of moi sture by the Karl Fischer method and this provided the impetus for

developing methodology for the determination of moisture in edible oils by FTIR

spectroscopy. Both FF A and moisture analyses are used to control many steps in the

overall oil refining process, such as deodorizer efficiency and soap adsorption, and to

evaluate storage conditions and oil stability. Thus these analyses determine important

quality and process control parameters which need to be determined accurately and

precisely. An examination of the literature regarding the current status of FFA and

moi sture analysis, indicates that there is still substantial room for improvement in terms

of accuracy, specificity and practicality as weIl as a need for methods that can be

automated by QC and process laboratories. Chapter 2 of this thesis reviews and details

the relevant analytical methodology and approaches to the analysis of FF A and moisture

to pro vide a context for the FTIR methods developed. In Chapter 3, a' method for the

/~. determination of FF A based on a stoichiometric reaction coupled with differential

spectroscopy is presented. Chapter 4 revisits the method, considering extraction as a

3

Page 21: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

,~

means of increasing sensitivity. Chapter 5 describes a method developed for moisture

analysis drawing in part on the successful approach used in Chapter 4 for FF A. Chapter 6

takes latter two methods and adapts and implements them on an auto-sampler equipped

FTIR to automate the methods attaining analysis rates of ~ 100 samples per hour as weIl

as testing them to determine their overall performance. Each chapter is presented as a

paper, Chapters 3-5 having been published, with Chapter 6 in preparation for submission.

It has been a challenge and learning experience to develop methodology which 1

personally feel will eventually impact the edible oil industry in a tangible way in the

future.

1.3. REFERENCES

1.

2.

3.

4.

5.

6.

Consumer expenditure on oils and fats: National statistical offices/OECD/EurostatlEuromonitor International

AOCS, Official Methods and Recornrnended Practices of the Arnerican Oil Chernists' Society, 4th edn., AOCS Press, Champaign, IL, 1998.

IUPAC, Standard Methods for the Analysis of Oils, Fats and Derivatives, 7th edn., Blackwell Scientific, London, 1992.

O'Brien, R. D., Fats and Oils: Forrnulating and Processing for Applications, Technomic Publishing Company, Lancaster, PA, 1998.

Laboratory Proficiency Pro gram, Cottonseed oil, AOCS, Champaign, IL, 2002.

van de Voort, F. R., Sedman, J. and Russin, T. Lipid Analysis by Vibrational Spectroscopy, Eur. J. Lipid Sci. Technol. 103,815-826 (2001).

4

Page 22: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

CHAPTER2

LITERATURE REVIEW

2.1. FREE FATTY ACIDS (FFA) ANALYSIS

2.1.1. Introduction

Free fatty acids (FF A) are common triglyceride hydrolysis products found in

crude oils and to sorne extent in refined oils as weIl as developing as a result of oxidation

or frying, ultimately impairing oil quality and functionality. ChemicaIly, FFAs are less

stable than triglycerides and therefore more likely to oxidize and cause ranci dit y (1).

From a quality standpoint, the FF A content of edible oils is not only a crucial factor

associated with their quality and economic value, but also an important quality indicator

in relation to fats and oils processing, i.e., serving as a measure of deodorizer efficiency

(1). The AOAC, AOCS, and IDPAC aIl have standard methods for the determination of

FF A, with the AOAC and AOCS methods being identical. These methods are based on

the titration of an oïl (3.5-56.4 g) dissolved in hot neutralized 95% ethanol (50-100 ml)

(2) or ethanol/diethyl ether (50-150 ml) (3) with a strong base to a phenolphthalein end

point under vigorous shaking. Although simple, these titrimetric methods are tedious and

consume substantial amounts of solvent which are costly and difficult to dispose of

safely; in addition, the use of a colorimetric indicator is problematic when dark crude oils

are analyzed.

In recent years, a variety of approaches have been investigated as possible

alternatives to the manual titrimetric methods employed to determine the FF A content of

oils. These generally require more sophisticated equipment such as flow injection

systems, automatic titrators and chromatographic instrumentation, among others. Many

of these techniques offer sorne benefits in terms of speed of analysis, amenability to

automation, and/or a reduction in the use of solvents, with the attendant environmental

benefits, and sorne provide a substantive gain in sensitivity over that attained with

titrimetric methods. The objective ofthis section is to review in general terms a variety of

approaches to FF A analysis and place the development of FTIR methodology in context.

5

Page 23: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

2.1.2. Historieal Perspective

After W orld War II, the food industry began a new stage of development and

expansion to catch up with the increased need for food as a result of the growing

population. As an integral part in the food chain, the oil industry expanded dramatically

which led to revisionlre-evaluation of many quality associated parameters and the ,

methods that are used to determine them. FF A determination was among the important

methods that were subjected to many studies and modifications to make it more sensitive

and reproducible as weIl as more rapid and economical. Over the past sixt Y years, many

methods were published, and each method utilized the analytical technology advances

associated with its time to improve the method as weIl as addressing the health and

environmental issues of concern at the time of publication. For example, as

spectrophotometry was the most weIl known analytical technique in the 1950s, most of

the FFA methods published at that time (and later) were colorimetric methods as opposed

to the previous generation of methods which were characterized by the use of manual

titration for quantification. Later on, the methods tended more towards electrochemical

techniques.

2.1.3. Colorimetrie Methods

Copper soap method. The first attempt to develop a colorimetric method for FF A

was carried out by Ayers in 1956 (4). FF As were converted into water-insoluble soaps by

reacting them with an aqueous solution of copper nitrate or cobalt nitrate. Quantification

was carried out by extracting the colored soap precipitate (blue for copper and pink for

cobalt) in chloroform followed by measurement of absorbencies at 675 nm for copper and

527 nm for cobalt. Although the method was not intended for FF A analysis per se, it

provided a basis for other researchers to develop colorimetric methods for FF A analysis

in edible oils.

In 1964, Baker (5) introduced his improved version of the cupric method which

was designed especially for FF A analysis in edible oils. The method involves dissolving

the oil sample (8 g) in benzene (50 mL) followed by a mixing step where an aliquot (10

mL) of the mixture is shaken with aqueous cupric acetate solution in a separate tube. The

mixture is centrifuged or let settle until a clear separation between the two phases is

6

Page 24: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

obtained. Quantification of FF A is determined by measuring absorption of the upper

benzene layer at 640 nm. A schematic diagram of the procedure is shown in Figure 2.1.

Although Baker claimed a good correlation between his method and the titrimetric

method, he indicated that certain copper soaps such as copper salts of palmitic and stearic

acids do not dissolve in benzene. Discrepancies between the parameters used by Barker

and Ayers, such as the working wavelength (640 vs. 675 nm), led Bains et al. (6) in the

same year to examine the spectra and the effect of oil color on the measurement. The

author concluded that 670 nm is the optimum wavelength, giving high sensitivity for

quantification of copper soaps in colored or non-colored oil.

jiP 11 ----11> .. ;:

Benzene + Où

,",

r---+-----, Aqueous Cupric Acetate '---_S_hak_e -', solution

~ Centlifuge 1

~

i JI Spectrophotometer

À= 640nm

Figure 2.1. Schematic diagram of the copper soap method for FF A analysis introduced by Baker in 1964 (5).

Despite the simplicity of the method, not much work was done to understand the

chemistry of the products formed or even to optimize the conditions of the method to

make it more sensitive and reproducible. In 1976, Lowry and Tinsley (7) published a

considerable work about the conditions and chemistry involved in order to provide a basis

to optimize the method and make it more sensitive and reproducible. The outcome of

7

Page 25: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

their work was a stabilization of the copper soaps of saturated FF A in benzene as weIl as

an increase in the sensitivity of the measurement by a factor of 10 just by controlling pH

by adding pyridine. Due to both health and practical concems, a number of publications

came out suggesting a replacement for benzene, including toluene (8), isooctane (9) and

·cyc1ohexane (10).

In 1981, Ekstrom and Rhee (8) introduced the first automated system for FF A

analyses in edible oils using the copper soap method which was able to analyze up to 20

samples/h. Automation was achieved using a flow injection analysis (FIA) system to mix

oil samples with the solvents and reagents used in the method. According to the setup of

the system, separation of the organic phase from the aqueous one was achieved online

through the use of segmentors and separators at specified locations in the system. In

contrast with the previous versions of the method, the measurement was done in the

aqueous phase instead of the organic phase. This was possible by implementing two

mixing-separation steps in the system; the first one was responsible for mixing the oil­

solvent mixture and cupric solution together to produce FF A-copper soaps and then

separate the two phases from each other (Figure 2.2). Wasle

Waste

-e- Pump l:x:xl MixingCoit

I:§J Injection Valve ~ Separator

• Segmentor

~~ Figure 2.2. Flow diagram of the FIA instrument for determination of FF A in oil samples as proposed by Ekstrom and Rhee (8).

8

Page 26: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

The second step involved extracting the copper ions in FF A-copper soaps into

another aqueous solution containing EDTA; and then the copper-EDT A complex formed

was measured spectrophotometrically at 660 nm. The system did not receive much

attention, mainly because it was complicated and had stability problems. Six years later,

Canham and Pacey (11) reported an improved FIA system using a one-extraction-step

process and improved hardware (segmenter and separator) illustrated in Figure 2.3. The

new system had improved stability and shorter analysis time (130 samples/h), but at the

expense of sensitivity. The sensitivity of FIA systems was boosted by Puchades et al. in

1994 (12) mainly by optimizing operating conditions, such as flow rate, pH, tubing

length/diameter, and oil/solvent proportion, while maintaining good analysis speed (up to

75 samples/h)' Two years later, Zhi et al. (13) came up with a new solvent extraction FIA

(SEFIA) setup and protocol for FF A analysis without requiring a segmenter or separator.

The method had the same high sensitivity as the previous method, but the analysis time

was significantly increased (12 samples/h) and the FIA system required substantial

coordination and timing which was provided by a minicomputer.

Waste

-®- Pump 1 :::>c:>c 1 Mixing Co.il

t:§:l Injection Valve Restriction Coi1

• Segmentor Separator

Figure 2.3. Flow diagram of the FIA instrument for determination of FF A in oil samples as proposed by Canham (11).

9

Page 27: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

Despite the advantages offered by the copper soap method (limit of detection

~0.01 %FFA) and its automated versions (~12-130 samples/h), the method did not

receive much attention from the oïl industry, mainly because of the complexity of the

analysis compared to the official method. Moreover, the FIA systems employed in the

analysis had specialized hardware (segmenters and separators) which made them not

suitable for other analyses.

Phenolphthalein based method Despite the drawbacks of the "official" method,

its simplicity and wide recognition made its automation a necessity in order to shorten the

analysis time and reduce the ambiguity in determining the end point. The first substantive

attempt to automate the official method was carried out by Linarers et al. in 1989 (14).

The setup of the system was simple (Figure 2.4) and used the same reagents as the

IUP AC official method (3) whilst quantification was based on monitoring the decrease in

the phenolphthalein intensity (at 562 nm) as a result of the reaction between KOH and

FF A in the oil sample. The system had an analytical range of 0.15-0.81 % FF A and

sample throughput of about 60 samples/h. In 1997, Nouros et al. (15) reported a

modification with a similar analytical range, the mixing coil being replaced with a mixing

chamber plus using a less expensive and less problematic solvent (l-propanol) as a

diluent and carrier. The sampling rate was reported as 30-100 samples/h, and solvent

consumption as 3-7 ml per sample. A similar system equipped with a photo probe

connected to a spectrophotometer was reported in 2001 by Mariotti and Mascini (16) for

FF A measurement in virgin olive oil. The system had the same sensitivity as the previous

ones with a sample throughput of 12-60 samples/h.

Phenol red with reverse micelles. In 1990, Walde (17,18) developed a new

approach for FF A determination in non-aqueous medium without titration. The method

relies on the acidity of FF A to cause color changes to the alkaline form of phenol red.

The indicator, which is water soluble, is dispersed in the non-aqueous medium

(oil/isooctane) by bis(2-ethylhexyl) sodium sulfosuccinate, a surfactant. Quantification

was achieved through monitoring changes in absorbance at 560 nm (disappearance of the

red color). Although this method was economical in oil and solvent usage (0.1 ml and 5

ml, respectively), it was found to be lacking in accuracy and no further attempts have

been made to optimize or automate it.

10

Page 28: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

Waste

-®- Pump 1 )OC 1 Mixing Coil

~ Injection Valve

Figure 2.4. Flow diagram of the FIA instrument for determination of FF A in oïl samples as proposed by Linarers et al. in 1989 (14).

2.1.4. Electrochemical Methods

Voltammetric techniques. Voltammetric techniques are widely used for

compounds that are readily reduced or oxidized to provide highly selective and sensitive

quantitative methods. Although FF A can be oxidized or reduced, they are by nature

weakly active electrochemically, a factor which impeded the development of reliable

electrochemical methods for FF A determination. The situation was changed in 1972

when Takamura et al. (19) published their work on the voltammetric behavior of quinone

in the presence of acids in unbuffered protic solvents. It was reported that addition of

acids to the reduced form of quinone (dissolved in ethanol) gave a very strong

voltammetric signal. The behavior of quinone has been utilized by many authors (20-25)

to develop FIA and HPLC methods for FF A analysis in edible oils. The FIA system

coupled with voltammetric detection had high sensitivity/reproducibility (0.05 % FF A,

RSD 0.7% (23», low solvent consumption and outstanding throughput (60 samples/h)

~', pH-Metric methods. The pH meter obviously cornes to mind as a means to

measure acidity (FF A); however, the high viscosity and low polarity of oils have made it

11

Page 29: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

impractical to use pH-metric methods. In 1991, Labshina et al. (26) suggested extraction

of FF A from oïl samples into a polar solvent to carry out pH measurements. In order to

facilitate extraction and pH measurement, the authors suggested use of a weak base,

triethanolamine (TEA), in a suitable solvent mixture (water (1 %), diethyl ether (80%) and

chloroform (19%» to convert carboxylic acids into their salts. Although workable, the

method had poor reproducibility, mainly because of instability of the electrode and the

basic reagent in the solvent mixture (due to the low concentration ofwater). The method

was later modified by Tur'yan et al. (27), who replaced the former solvent mixture with a

1: 1 (v:v) mixture of iso-propanol and water containing 0.2 M TEA and 0.02 M KN03, the

latter added to obtain a constant ionic strength. Rather than extracting FF As, this solvent

mixture effectively makes an emulsion within which the pH measurement is made.

Various versions of this approach were developed for acidity measurements in petroleum

oils (28), oilseeds (29) and hydraulic oils (30). Although providing a simple means of

measuring FF As, it does not provide any improvement over the official titrimetric method

in terms of sensitivity or solvent consumption.

Automated potentiometric titration. One of the main reasons that the official

method is still in use is the advances that have been made in potentiometric

measurements and automated titration systems. The first application of potentiometric

measurements to determine the end point of acidlbase reactions goes back to the early

years of the last century. It was realized that a potentiometric end point is sharper and

more reproducible in determining the end point of a reaction than using a color indicator;

however, a lack of the technology required to automate the measurement limited the

acceptance of this approach as a practical alternative to the conventional titration. Since

W orld War II, substantial effort has been devoted to improving potentiometric response

as well as automation and has resulted in the commercial availability of potentiometric

auto-titration systems (31-35). The basic setup of such a system consists of a

potentiometric electrode (pH meter), syringe pump, and stirrer and a recorder (or

computer). Initial automated titrators were large and had few features, but

computerization has overcome many of their initial limitations, resulting in general

acceptance in analytical labs. Figure 2.5 presents a modem potentiometric auto-titration

system which in brief, involves volumetric addition of the titrant (alkaline solution) from

12

Page 30: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

the bottle at the rear (see Figure 2.5) in small increments by the syringe pump while the

pH is measured by the electrodes immersed in the beaker of analyte on the stirrer. By

convention, the titration curve has the volume of titrant added on the x-axis and the

corresponding change in pH (expressed as pH units or m V) on the y-axis. The end point

of the titration is determined as the midpoint of the maximum slope of the titration curve.

Once the end point is detected, the volume of the titrant added stoichiometrically defines

the concentration of FF A in the oil sample.

Figure 2.5. Metrohm auto-titrator instrument equipped with autosampler for FF A analysis.

Auto-titrators are available comlJlercially from a number of suppliers, such as

Metrohm, Mettler, Labcor, etc., and can be equipped with many optional accessories, e.g.

autosampler, different sens ors, etc. For FF A analysis, most of the titrators use glass

electrodes to measure changes in pH, al: 1 mixture of ether and etllanol as diluent, and

13

Page 31: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

KOH/ethanol (or KOHltert-butanol) (36) as titrant. The accuracy of the auto-titrator

systems depends mainly on the capability of delivering small and consistent amounts of

the titrant, which varies between different manufacturers. The sensitivity, on the other

hand, depends more on the pH electrode; if the electrode receives proper conditioning

treatments, good sensitivity (on the order of 0.01% FFA) will be maintained.

Potentiometric auto-titrator apparatuses are the most widely used systems for FF A

analysis in quality controllabs.

2.1.5. Thermometrie Methods

Thermometric methods are titrimetric methods that use the change of temperature

resulting from a chemical reaction to determine the end point of the titration. The

principle of this type of method is the same as that for potentiometric auto-titration; the

only difference is the type of sensor (detector). The detectors (usually thermistors) used

in this kind of method show large changes in their electrical resistance for small changes

in temperature. The first considerable advance in determining the end point of the

acid/base reaction in non-aqueous medium thermometrically dates back to the 1960s. In

1965, Vaugham (37) reported that upon titration of a sample containing a weak acid

dissolved in acetone, the first excess of base catalyzes a strong exothermic reaction which

can be detected easily. Since the introduction of this method, many modifications have

been made which involve the use of different reagents and solvents in order to make the

method more sensitive and the end point sharper. In 2003 Thomas (38) reported that

acetone and chloroform react exothermically (and violently under certain conditions) in

presence of base, and performance experiments using a commercial automated

thermometric titration unit with KOH in iso-propanol as titrant showed exceptional

reproducibility in determination of FF A in edible oils. Titrations using thermistor

detectors are becoming popular in quality control labs, because these sens ors are largely

maintenance-free.

2.1.6. Current Status ofFFA Analysis

Many of the instrumental methods discussed above offer substantial benefits over

traditional manual titration in terms of speed of analysis, amenability to automation,

and/or a reduction in the use of solvents or amelioration of disposaI costs. However, the

14

Page 32: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

colorimetric methods have inadequate sensitivity, with the additional drawback of

requiring frequent calibration, which makes them generally impractical for routine

analysis. Electrochemical methods require frequent titrant standardization, and the high

accuracy and reproducibility that these methods can provide are also dependent on the

conditions of the electrodes, making regular maintenance necessary. Moreover, none of

the methods discussed pro vide specificity to FF A measurement as they tend to be based

on measures of total acidity rather than FF A per se. Thus, there remains a need for a

specific, accurate and simple method for the analysis of FF A in edible oils.

2.1. 7. Infrared Spectroscopie Methods

The possibility of employing FTIR spectroscopy in FF A analysis has been

investigated by a number of researchers over the past fifteen years. The first such study,

reported by Lanser et al. (39) in 1991, was directed toward the development of a semi­

quantitative method for the analysis of soybean oils containing appreciable levels of FF A.

The FTIR spectra of oils placed between two KBr windows were collected, and the

spectral changes observed in the carbonyl region, after deconvolution of the spectra over

the 2000-1600 cm- I range, were correlated to the FFA content obtained by the standard

AOCS method. The method was placed on a quantitative basis by devising a calibration

using standards prepared by spiking oleic acid into FFA-free soybean oil. It was noted

that because the carboxylic acid c=o absorptions appear as a broad shoulder on the

strong triglyceride ester linkage absorption, a major limitation to making the method

more quantitative was the spectral variability of the ester band among different oils. In

1993, Ismail et al. (40) developed a more rigorous quantitative method based on

measuring the carboxylic acid c=o band at 1711 cm- I after applying oil samples in their

neat form onto an attenuated total reflectance (ATR) cell. The matrix effects noted by

Lanser et al. (39) were reduced by ratioing the calibration and sample spectra against the

spectrum of an FF A-free oil of the same type as the oil being analyzed, rather than an air

background spectrum. In recognition of the limited utility of this approach in eliminating

matrix effects in the case of crude and oxidized oils (40), an alternative indirect method

was also developed. This method involved adding KOHlMeOH to the oil so as to convert

any FFAs present to their salts, effectively moving the measurement band from 1711 cm- I

to 1570 cm-l, a region where there are no major interferences which limit the accuracy of

15

Page 33: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

peak height measurements (see Figure 2.6). In addition to overcoming matrix effects, this

indirect method was more sensitive than the direct measurement of FF A absorption,

albeit at the expense of complicating the analysis. However, the method was susceptible

to error due to saponification of the oil by the KOHlMeOH reagent.

0.3

0.2 ~ A B t:J

== ~ ,.Q .... 0 0.1 r7J

,.Q

<

0.0

-0.14-----------~------------~----------~ 1800 1700 1600 1500

Wavenumber (cm"1)

Figure 2.6. Shift in FF A band after conversion to carboxylate salt. FF A shows band at 1711 cm- l (A); the band shifts to 1570 cm- l after addition of base (B). (The spectra were taken in our lab)

After these initial studies, two publications dealt with FF A analysis by FTIR

spectroscopy in relation to palm and olive oil, respectively (41,42), both using partial

least squares (PLS) to develop relationships between spectral changes and the results

obtained with the standard method. Subsequently, Verleyen et al. (43) followed up more

rigorously on the initial work of Lanser et al using peak height measurements at 1711 cm-

1 to develop workable calibrations for a variety of oils but ultimately concluded that the

calibrations were heavily oil dependent.

The most sophisticated work related to FF A analysis has been that of Caiiada et

al. (44), who developed an elegant automated FTIR-based continuous-flow analysis

16

Page 34: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

system capable of analyzing -40 samples/h. For this system, these authors adapted the

indirect KOH/MeOH method developed by Ismail et al. (40) as an effective means of

overcoming the matrix effects associated with direct measurement of the carboxylic acid

C=O band. Having noted the susceptibility of oils to saponification by KOH even within

the 30-s mixing time employed, Caiiada et al. modified the system to eliminate this

source of error by reducing the contact time between KOH and the oil (premixed with

MeOH) to -2 s.

Aside from these FTIR studies, FF A analysis by other types of vibrational

spectroscopy, namely, near-infrared (NIR) and Raman spectroscopy, has also been

investigated (45-49). AlI the NIR and Raman methods that have been published to date

are direct methods in the sense that there is no sample preparation involved and the FF A

measurement is done using the spectrum of the neat oil. In addition, all these methods use

PLS and/or multiple linear regression (MLR) calibrations that have been developed for

specific oil types and hence cannot be applied to other oil types or samples of unknown

origin. Moreover, the NIR and Raman methods, although having the benefit or being

reagent-free, tend to have relatively low sensitivity (~ 0.3%), which makes them suitable

largely for the analysis of oils known to contain high amounts of FF A such as fish

(45,46), palm (47), and olive oils (48, 49).

2.2. MOISTURE CONTENT ANAL YSIS f '1.

, . 2.2.1. Introduction

Moisture content in fats and oils is considered a key quality parameter in the

edible oil industry and has to be controlled throughout oil processing, this being

especially important when oils are being pressed from high-moisture raw materials.

Moreover, water is commonly added to oils during degumming and refining operations

which then needs to be removed, usually by centrifugation and/or vacuum drying.

Attaining a low moi sture content after refining is very important for the effective

adsorption and removal of traces of soap in the oil (2).

17

Page 35: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

Water has limited solubility in oils and fats, ranging from 0.05 to 0.3% (50);

however, from a quality standpoint, the moi sture content should be reduced to the lowest

practical level. By convention, moisture content of dried refined oil should be below

0.1% and is most often on the order of 0.05% (1). Hydrolysis, breakdown of fat, is

induced by the presence of moi sture, accelerated by heat and/or residual enzymes, and

results in the formation of FF A, di- and monoglycerides and glycerol. HYdrolysis can

result in off-flavors, a reduced smoke point, increased fat absorption and darkening of

shortenings used for frying (1). Maintaining a low moisture content williimit hydrolysis

of triglycerides during processing and storage (1) and avoid the formation of FF A, which

are susceptible to auto-oxidation. Thus, the routine determination of moisture would be

helpful in ensuring acceptable oil quality; however, moi sture analysis in oils is both

difficult and problematic.

There are many classic methods for the determination of moisture in edible oils

and related products, but they can be grouped into three categories: evaporation,

distillation and titration procedures. The Official Methods of the American Oil Chemists'

Society (51) lists several evaporation procedures which involve heating oil samples on a

hot plate (Ca 2b-38), in an oyen (Ca 2c-25), or in a vacuum oyen (Ca 2d-25). Although

these procedures are simple, there are many problems associated with them. For example,

the evaporation of moi sture may also remove other volatile components, such as short­

chain FF As, and as a result the weight loss recorded is not only due to loss of moisture.

Highly unsaturated oils can oxidize d~~g the evaporation process and an increase in . . .

weight may actually be recorded, while triglycerides may undergo hydrolysis by the

moi sture initially present giving rise to FF As.

The distillation approach is based on the azeotropic property ofwater which is the

ability of water to form constant boiling mixtures with many organic solvents in which it

is immiscible when cold. The AOCS distillation method (Ca 2a-45) involves mixing 20-

200 g of the test sample with 100-300 ml oftoluene. The mixture is heated and water is

distilled out of the sample as a constant boiling mixture with the solvent. The distillate

separates with water forming the lower phase, whereupon its volume is measured in a

specially graduated tube. Although it is an accurate procedure per se, the distillation

method has many drawbacks that limit its use, such as its limited sensitivity (not

18

Page 36: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

applicable to samples containing less than 0.5% moisture), the large sample size required,

and its lack of suitability for high-volume testing.

The "best" method available for moisture determination is the Karl Fischer

titrimetric method (e.g., AOCS Ca 2e-84), and it is the gold standard against which

proposed new methods for moi sture analysis are compared. As such, this method will be

discussed in detail, and proposed modifications that have been reported in the literature

will be surveyed to provide a c1ear picture of its current status.

2.2.2. Karl Fischer Method

The Karl Fischer (KF) method involves the chemical reaction of water with the

Karl Fischer reagent (mixture of iodine, sulfur dioxide, pyridine (or similar base) and

methanol), which consumes water stoichiometrically. The general reaction is as follows:

S02 + CH30H + B ( ) CH3S03 (HB)

12 ,1;+ H 20 + CH3S03(HB) + 2B~CH3S04(HB) + 2(HB)1 [2.1]

where B is a base, such as pyridine or imidazole. The general protocol involves mixing

5-25 g of oïl with anhydrous methanol (or 1-propanol) and titrating with Karl Fischer

reagent to a cherry-red colored end point. Due to the high error associated with manual

titration (0.6% relative error) (51), the,titratiop is frequently carried out automatically in a 1 i ,

"

c10sed vessel where the end point is determined potentiometrically. The automated

method can determine as low as 200 ppm of water. Although detection of the end point

by potentiometric measurement is more sensitive than in the case of manual titration, it is

still a titrimetric method and, as such, standardized reagents are required to obtain

consistent and accurate results. The need for standardization is eliminated by using

coulometry; an additional advantage of coulometric measurements is their higher

sensitivity, which in the case of the KF reaction extended the limit of detection to coyer

the range 1-25,000 ppm water.

Although the KF reagent is specific to water, the presence of aldehydes and

ketones can complicate the analysis by introducing side reactions. Two side reactions are

19

Page 37: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

weIl known to take place during KF measurement which directly interfere with accurate

moi sture determination and are believed to occur "simultaneously" in the presence of

aldehydes/ketones. The first reaction is acetal/ketal formation where methanol reacts with

aldehydes/ketones to produce water molecules as shown in Eq. 2.2.

H H, /OCH 3

\=0 + 2 CH 30H .. + H 20 C

RI R/ "aCH 3

R R, /OCH 3

)=0 + 2 CH 30H .. + H 20 C

R/ "aCH 3 [2.2] R

The occurrence of such reactions increases the number of water molecules in the

system, leading to an overestimation of moi sture content in the sample being analyzed.

The second reaction is called bisulfite addition. Bislfite ions are common products in KF

systems, being formed reversibly by the reaction between water molecules and sulfur

dioxide under basic conditions (Eq. 2.3). In the presence of aldehydes/ketones, the bislfite

ions undergo addition reactions and form bisulfite-addition products as shown in the

reaction scheme below:

H \ e ® c=o + HS0 3 + HB

1 R

\=0 + soie + 2 HB®

1 [2.3]

20

Page 38: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

As can be noted from these reaction equations, these reversible reactions consume

water, which is then released back into the system owing to the shifting of the

equilibrium to the left as water reacts with the KF reagent. Accordingly, these reactions

delay reaching of the end point, leading to an underestimation of the water content in the

sample when these reactions are not taken into account in setting the instrument

parameters.

The reaction conditions and rates of the above reactions were extensively studied

by many authors in order to minimize the interference as much as possible. In this regard,

Scholz (52) reported after studying the behavior of 44 difÎerent aldehydes and ketones

with different KF formulations that acetal/ketal formation reactions are acid-catalyzed

and that higher pH had a significant effect in reducing the extent of these side reactions.

He suggested use of the stronger base imidazole as a substitute for pyridine in the KF

. formulation in order to make the reaction conditions more basic. Scholz (52) agreed with

other authors who found that replacement of methanol with other solvents, such as 2-

methoxyethanol (53), dimethylformamide (54), and propylene carbonate (55), suppressed

acetal/ketal formation reactions, but he found that the bisulfite addition reaction is more

favored in nonalcoholic solvents; consequently, he suggested use of alcoholic solvents

and higher pH rather than use of nonalcoholic solvents as a means of suppressing side

reactions of aldehydes and ketones with KF reagents. Examination of different types of

alcoholic solvents (primary, secondary and tertiary alcohols) as an alternative to methanol

led Scholz to strongly recommend that f7cl1loroethanol and 2,2,2-trifluoroethanol be used

in the KF formulation instead of methanol. He reported that KF formulations using either

of these solvents with imidazole as the base had outstanding performance in

determination of water in the presence of aldehydes and ketones; in the same year he

patented a new KF reagent using those two solvents and imidazole (56). In 1987,

Andersson and Cedergren (57) reported that the bisulfite addition reaction is a slow

reaction and, therefore, suggested use of fast-reacting KF reagents or higher iodine

concentration to reduce the analysis time in order to make the onset of the bisulfite

addition reaction negligible. In addition, the authors (57) found no observable

improvements when 2,2,2-trifluoroethanol was used in place of 2-methoxyethanol. Bizot

(58) and Cedergren (59) reported that addition of sorne chemicals such as formamide to

21

Page 39: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

pyridine-buffered methanolic KF reagent was found to speed up the KF reaction by a

factor of 100, which in turn can minimize the problems connected with slow reaction

rates between KF reagents and water. Altematively, Scholz (60) reported that

replacement of pyridine with imidazole (more basic than pyridine) was found to increase

the speed of the reaction of the KF reagent with water and ultimately minimize the effect

of both the bisulfite addition reaction and acetal/ketal formation. In 1994, Oradd and

Cedergren (61) showed that in order for imidazole to speed up the reaction the pH has to

be in the range 7-10.

Besides the side reactions of water productionlconsumption due to presence of

active carbonyl groups, there are other important side reactions associated with iodine. In

coulometric KF analysis, iodine is formed at the anode and reacts quantitatively with

water present in the system as shown in Eq. 2.1, with the end point of the titration being

the point at which iodine is no longer consumed. Iodine is known to be a strong oxidizing

agent and can react with substances such as thiosulfite, thiosulfate, ascorbic aeid,

hydrazines, hydroxylamines, Tl, Sn2+, In+ and Cu+, etc., in the matrix being analyzed,

resulting in overestimation of the moi sture content. On the other hand, certain oxidizing

agents such as Cu2+, Fe3+, N02", Br2, Ch and quinines would oxidize r (one of the KF

reaction products) to h, and thus the presence of these species in the matrix would lead to

underestimation of the moi sture content. In addition, h can combine with r to pro duce

the triiodide ion (h):

l , . i 1°'" 1" 2;h-,~ 3 [2.4]

Although this ion will eventually react with water, it has been reported that reactions

involving the triiodide ion are about 4-fold slower than those of iodine itself (62), which

may lead to underestimation of water content. To minimize the above-mentioned

problems, it is strongly suggested to use fast-reacting KF reagents combined with

optimum conditions in order to speed up the reaction so that less side reactions will have

chance to occur (63,64).

Another important modification of the KF method was that introduced to make it

suitable for the determination of water in hydrophobic liquids such as edible oils and

lubricants, because the KF reagents as originally formulated (65) were not miscible with

22

Page 40: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

such samples. It was suggested to modify the polarity of the KF working medium by

adding solvents such as chloroform (66), xylene, hexanol, I-propanol, l-octanol (67),

propylene glycol and others. Partial solubilization of oil samples in KF reagents was

found to be one of the main reasons for the commonly observed variation in results

obtained by the KF method (68).

Most of the progress that has been made in understanding and improving KF

reactions has been adopted commercially. An important consideration in making

commercial formulations is the toxicity of the reagents, and the pyridine employed in the

original KF reagent has largely been replaced by the less noxious imidazole, which also

serves to speed up the KF reaction, as discussed above. Different formulations of KF

reagents are available under various brand names, the most common being

HYDRANAL ®, representing pyridine-free reagents, which are produced in different

formulations to suit a variety of applications. On the other hand, this complicates the

univers al comparison of KF results as the results obtained depend on the formulation

used.

KF reaction cell. In conjunction with the work done with reagent formulations to

make the KF method faster and less susceptible to side reactions, substantial development

has taken place in relation to the design of the reaction vesse!. The conventional

coulometric cell inc1udes a potentiometric detector, a cathodic and an anodic

compartment, and a diaphragm. The cathode is in electrolytic contact with the anode

through the diaphragril, whi~h pl~ysa:v:éry·important role in restricting the iodine that is

being produced at the anode from reaching the cathode, where it can be reduced by

reaction products such as thiosulfate, hydrogen sulfide and other products that are

believed to be produced at the cathode (69,70). Optimization of the dimensions of the

cathode and anode and the possibility of removing the diaphragm have been extensively

studied to make the design simpler and facilitate maintenance. As a result, many designs

have been proposed with various cathode dimensions, with or without a diaphragm, and

with the use of constant or pulsed currents (71-79). The use of a diaphragm eliminates the

possibility of iodine being reduced by products at the cathode and allows the use of

electrolytes other than the KF reagent in the cathodic compartment of the cell. Clogging

of the diaphragm, however, is a common problem and the cell requires long conditioning

23

Page 41: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

times before startup (~2.5 hl. On the other hand, diaphragm-free systems have shorter

conditioning times and are easier to maintain, but they suffer from the formation of

oxidizable reduction products at the cathode. This problem has been minimized

considerably by implementing rapidly reacting reagents combined with optimizing the

dimensions of the cathode and the current density. The latter were found to be critical

factors in determining the analytical accuracy obtained with the coulometric cell (72, 77-

80). The type of current used in the coulometric ceIl, pulsed or constant, was found to be

critical, especially for diaphragm-free cells, in reducing the rate of development of

oxidizable products at the cathode and also to minimize the drift at the end point. In this

regard, pulsed current was found to be the most suitable for diaphragm-free cells (78-80).

Investigation of such cells also indicated that the titration rate is a critical parameter in

relation to analytical accuracy (79), but the optimum varies from one system to another.

KF detectors. Many end point detectors have been used with the KF method.

Determining which detector is best from the literature is not easy because comparisons

are not straightforward as the end-point detection is influenced by a number of factors.

These include the mode of titration (volumetric or coulometric), cell design, electrode

response, the extent of background drift caused by moisture diffusion, side reactions and

the use of different reagents and additives in the titration.

Early on, amperometric detectors were commonly in use, largely supplanted later

(81,82) by controlled-current potentiometers (bipoteniometry) (59, 83-86). In general,

both these types of detectol'S use a' consùmt current between two platinum electrodes

while measuring the voltage needed to sustain the current. At the equivalence point of the

titration, there is a sharp drop in the voltage required to sustain the current because excess

12 is present and the current can be carried at a very low voltage in its presence. Although

controlled-current potentiometers have better accuracy and shorter analysis times than

amperometric detectors, their accuracy was found to be affected by both the electrode

dimensions and the external current generated between the cathode and the anode.

Recently, zero-current potentiometers (true potentiometry) have received substantial

interest (66,71,87,88) because they offer better selectivity and do not suffer from the

electric field interference generated in the coulometric cell. Although it is the detector of

choice in modem KF coulometric ceIls, this type of detector must meet certain

24

Page 42: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

requirements to perform well such as a suitable reference electrode system and a

calibration procedure to establish the relationship between the redox potential and the

concentration of excess iodine (87).

Spectrophotometric end-point detection. Upon addition of the KF reagent to a

sample, the original dark brown color due to the presence of iodine changes toward light

yellow depending on the amount of water in the sample that has reacted with the KF

reagent. This change in color implies that the development of a colorimetric method for

the KF method would be possible; however, very little work has been done in this regard.

The earliest attempt to employ colorimetric detection in KF titration involved measuring

the absorbance of the sample at 525 nm (89). This approach allowed detection of 100

ppm water with a precision of ±3 ppm. Subsequently, Dahms (90) patented a colorimetric

method for determination of moi sture using the KF reaction based on the injection of the

sample into a fixed volume of KF reagent and colorimetric measurement of the change in

optical density (OD) at 520 nm. The amount ofwater in thesample was determined from

a calibration curve established by measuring the OD of the reagent mixed with defined

amounts of water. Although the author used 520 nm as a working wavelength, he noted

that the KF reagent absorbs over a wide spectral range (500-620 nm) and the sensitivity

of the method depended on the wavelength used. This approach was adapted for use with

a flow injection analysis (FIA) system (91) and improved to correct for factors affecting

OD measurements such as dilution effects, the transmission characteristics of the cuvette,

the refractive index of the sample, e~co'" through the use of a reference dye that has an "Ii f

absorption (600 nm) far from the working wavelength (set at 420 nm) (92). A system

based on this approach combined with specially designed ready-to-use sealed vials

containing the KF reagent was developed (93,94). Although this system appears simple to

use, it has not been independently validated.

Flow injection analysis. Due to the many interferences affecting KF titrimetric

methods, various approaches have been explored to carry out KF analysis. As indicated

earlier, many authors suggested speeding up the KF reaction as a means to both overcome

the problems of side reactions and increase accuracy by minimizing the problems

inherent to coulometric analysis. These issues become even more important when one

considers FIA as a means of automating KF moi sture determinations. In 1980, Kagevall

25

Page 43: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

et al. (91) developed an FIA system for determination of moi sture in organic solvents

using both potentiometric and spectrophotometric detectors. The method used a fast­

reacting one-component KF reagent (formamide/pyridine buffer instead of pyridine

alone) which was diluted with methanol as a carrier solvent. The system was capable of

analyzing up to 120 samples/h and covered a range of 100-50,000 ppm moisture. The

authors noted that the potentiometric detector gave a better SD (:::;0.5%) than that

obtained spectrophotometrically. The efficacy of the system in minimizing side reactions

in iodine-consuming samples was tested using penicillin as the sample (95), with the

results indicating a considerable decrease in side reactions. A similar, but improved FIA

system was introduced by Escott and Taylor (96) to analyze gasoline-alcohol blends. A

methanol-xylene mixture was used to dilute samples and make them miscible with the

methanolic KF reagent. Quantification was based on direct potentiometric measurement

and the system, which was equipped with an autosampler, was able to analyze up to 60

samples/h over a range of 1-1500 ppm moisture. The results tended to be higher than

those obtained using a titrator, but the reproducibility was better. The variables affecting

the performance of the KF method in the FIA mode using potentiometric or

spectrophotometric detectors have been thoroughly investigated by Dantan et al. (97).

The authors suggested new FIA designs for both types of detectors to pro vide more

flexibility in reagent concentrations as well as on-line analysis capability. The optimized

methods permit rapid, precise and automated determinations of moi sture in a wide range

of samples over the range of 100-50,000 ppm with a reproducibility of better than 3% • l, j

RSD. According to the authors, alteraÙon of~ertain parameters, such as injection volume,

detection wavelength and/or the concentration of KF reagent, could further extend the

analytical range, with the presence of traces of water in the carrier solvent being the main

factor determining the detection limits of these methods.

Current status of the KF method. As can be seen from the above discussion, most

of the modifications of the KF method proposed to date have focused on minimizing the

inherent problems associated with the KF reaction itself. These modifications have

minimized the problems of side reactions and increased the diversity of samples that can

be analyzed, but at the cost of complicating the method due to the introduction of

different KF reagents and apparatus designs. The modifications that were introduced over

26

Page 44: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

the years made the method suitable for the analysis of almost an types of samples;

however, ultimately the accuracy of the method is dependent on selecting the right

combinations of a long list of variables and operating parameters. The selection of

variables is difficult because it requires not only knowing what is in the sample that is

being analyzed but also understanding clearly the chemistry involved. In this regard,

Margolis (68,98) noted that systematic biases were observed in interlaboratory

collaborative studies and depended on both the type of apparatus and the nature of the

solvent system employed in the measurement. He tested a number of hydrocarbon-based

lubricating oils using different KF methods (volumetric and coulometric) and different

analysis conditions (different solvent compositions and different instrumental parameters)

and concluded that the measurements were very much dependent on the nature and

concentration of the non-polar organic solvent and the miscibility of the oil with the KF

reagent. The severity of the bias can be minimized by selecting the appropriate working

parameters and reagent composition for the sample at hand, but these are not readily

evident in many cases. Recently, Larsson and Cedergren (99) studied factors influencing

the accuracy and precision of moisture determination in oil using a diaphragm-free KF

coulometric cell using eight different types of commercial coulometric reagents and

various modifications. The best results were obtained with a homemade coulometric cell

with fully adjustable parameters and reagent formulations not commercially available.

Although one of the authors (Cedergren) has spent more than 20 years working with the

KF method, he declared that there is more research needed to better understand the

function of the cathode reaction in the diaphragm-free coulometric cell when using

different types of KF media. Aside from these issues, the coulometric cell has the

disadvantage of requiring regular maintenance (weekly) and long conditioning times.

Even with its many problems, the KF method remains the gold standard for moi sture

determination as no better alternative has been developed.

2.2.3. Infrared Methods

At first glance, FTIR spectroscopy would appear to be a viable technique for the

measurement of moisture in oils, as water absorbs strongly in the IR portion of the

spectrum (Figure 2.7), exhibiting intense and readily measurable bands in the regions of

27

Page 45: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

3700-3000 and 1670-1630 cm- l due to its O-H stretching and H-O-H bending vibrations,

respectively (100).

0.4

0.3

~ CJ

= c: 0.2 ,.Q "-0 C7.l

,.Q

< 0.1

o.o~---

4000 3500 3000 2500 2000 1500 1000

Wavenuniber (cm-1)

Figure 2.7. FTIR spectrum ofpure water.

However, FTIR moisture analysis in oils is complicated by spectral interferences

from other OH-containing constituents, such as FF A, mono- and diglycerides, alcohols,

and hydroperoxides, which may be present in substantial amounts (in terms of their

impact on moisture quantification both in crude oils and in refined oils that have

undergone auto-oxidation to any significant extent). Quantification of moisture is further

confounded by the hydrogen bonding interactions between water and the se species as

, weIl as between water and carbonyl-containing secondary oxidation products (aldehydes

and ketones), as such interactions affect the positions, band shapes, and intensities of the

water absorptions. Chemometric techniques such as PLS can, in principle, be employed

to compensate for these types of "matrix effects" but, in practice, their utility is limited by

the difficulty of meeting the criterion that aIl possible sources of spectral variability must

28

Page 46: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

be represented within the training set used to develop a calibration model. This limitation

is well illustrated by previous work conducted by the McGill IR Group on the

determination ofhydroperoxides in edible oils based on their Q-H stretching band at 3444

cm-l, which is subject to effectively the same confounding effects as moi sture analysis

(101,102). Recently, Che Man and Mirghani (50) published a PLS-based FTIR method

for the determination of moisture in crude palm oïl based on a set of calibration standards

prepared by spiking oïls with water to coyer a range of 0-13% moi sture content. The

method is simple to implement, involving recording the spectra of neat oil samples in a

1 OO-~m transmission cell and using the PLS algorithm to predict the moi sture content.

However, validation of this method by predicting an independent test set was not

performed, and hence there is no evidence that unknowns would be modeled adequately

by the limited number of spiked oïl samples used to develop the PLS calibration. Indeed,

the physical stability of the calibration standards, and hence the validity of the calibration

model, is questionable since phase separation of oïl standards having moisture contents as

high as 13% would seem likely. Apart from these issues, this work does not serve as the

basis for a global method, as the calibration model developed is specific for crude palm

oïl and does not take into account the complicating factors outlined above that must be

considered in the development of a generalized FTIR method for the determination of

moisture in edible oils.

A completely different approach to overcoming matrix effects was taken by van

de Voort et al. (103) for the FTIR dete~ination of moi sture in new and used mineral­

based lubricants, which are quite different matrices from edible oils but present similar

challenges from an FTIR analytical perspective. The approach invoives splitting the oïl

sample into two halves and treating one half with a dimethoxypropane (DMP)/dioxane

mixture to convert water present in the oïl into acetone via the reaction in Eq.2.5 prior to

recording its FTIR spectrum.

OCH3

1 CH3-C-CH3

1 OCH3

2,2-dimethoxypropane

® H

[2.5]

29

Page 47: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

The other half of the sample is simply diluted with dioxane, and its spectrum is

subtracted from that of the DMP-treated portion to generate a differential spectrum. The

amount of acetone that is produced in the reaction is quantified by measuring its v (C=O)

absorption at 1740 cm-1 in the differential spectrum and is related to the amount of water

present in the initial sample using a simple calibration equation. Owing to the high molar

absorptivity of this acetone band, very good sensitivity (±50 ppm water) was obtained

over the range 0-2000 ppm. Thus, the use of a reagent that reacts stoichiometrically with

water to produce a strongly IR-absorbing product followed by the application of

differential spectroscopy to isolate the spectral features associated with this conversion

provided a generalized method for the determination of moi sture in mineral-based

lubricants, without the complexities inherent in chemometric modeling of the sample

matrix. This approach, however, is not directly applicable to edible oils because the

triacylglycerol ester linkages give rise to an intense v (C=O) absorption that would

completely mask the v (C=O) band of acetone in a transmission-cell spectrum, except at

extremely short path lengths. Moreover, the use of ATR, which inherently provides a

short effective path length, to overcome this limitation is impractical owing to the

difficulty of preventing the sample from picking up moi sture from the environment or

losing the acetone produced, nor would it provide sufficient sensitivity for the analysis of

moisture at the low levels present in edible oils.

With regard to moisture analysis in edible oils by vibrational spectroscopic

techniques other than FTIR spectroscopy, a few papers on NIR methods have appeared in

the literature. The first paper, published in 1970 by Vornheder and Brabbs (104),

extended work published by other authors who determined water content in a variety of

samples (organic solvents (105) and food materials (106,107», but not edible oils. The

method involves mixing. oil samples with a polar solvent (DMSO) in order to extract

water molecules from the oil. The intensity of the water absorption band in the NIR

spectrum of the solvent is then measured in the range 2.1-1.8 J.lm (max absorption at 1.94

J.lm (5160 cm-1), and the water content is determined using a simple Beer's law equation.

Although the working range of the method depends on the sample:solvent ratio, the limit

of detection of the method was 300 ppm. Recently, Cozzolino et al. (46) used PLS to

30

Page 48: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

~"

establish a method for moi sture analysis in fish oil from the NIR spectra of neat oil

samples. The calibration equation had R2 of 0.8.

2.3. CONCLUSION

The standardized methods for the determination of both FF A and moi sture in

edible oils established in the middle of the twentieth century continue to be the basis for

the determination of these crucial oil quality parameters. Many modifications of these

methods or alternative techniques have been proposed, but it has been shown in the

discussion above that none of these have fully addressed the need for rapid, accurate and

simple instrumental methods that would be suited to automated process and quality

control. FTIR spectroscopy has the potential to meet this requirement, but this promise

has not yet been fulfilled in fairly extensive work to date on FF A analysis or in a single

study on moi sture analysis. Accordingly, the research described in the following chapters

of this thesis was undertaken to overcome the limitations of the FTIR methods developed

to date and allow the potential ofFTIR analysis to be realized.

2.4. REFERENCES

1. O'Brien, R. D., Fats and ails: Formulating and Processing for Applications, Technomic Publishing Company, Lancaster, PA, 1998.

2. Official Methods and Recommended Practices of the American ail Chemists 1

Society, 4th edn., Champaign, IL, 1989, AOCS Official Method Ca-5a-40.

3. Determination of the Free Fatty Acids in European Commission Regulation (EED) No. 2568/91, Official J01;1I'11al of the European Communities, No. L 248 (5.9.91), p. 6. " ,

4. Ayers, C. W. Estimation of the Higher Fatty Aeids C7-C18. Anal. Chim. Acta, 15, 77-83 (1956).

5. Baker, D. A Colorimetrie Method for Determining Free Fatty Acids in Vegetable Oils., J. Am. ail Chem. Soc., 41, 21-23 (1964).

6. Lowry, R. R. and Tinsley, 1. J. Rapid Colorimetrie Determination of Free Fatty Acids, J. Am. ail Chem. Soc., 53, 470-473 (1976).

7. Bains, G. S., Rao, S. V. and Bhatina, D. S. A Colorimetrie Procedure for the Estimation of fat Acidity in Peanuts and Peanut Meals. J. Am. ail Chem. Soc., 41, 831-832 (1964).

8. Ekstrom, L-G. An Automated Method for Determination of Free Fatty Acids. J. Am. ail Chem. Soc., 58, 835-838 (1981).

31

Page 49: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

9. Kwon, D. Y. and Rhee, 1. S. A Simple and Rapid Colorimetric Method for Determination of Free Fatty Acids for Lipase Assay. J. Am. Oil Chem. Soc., 63, 89-92 (1986).

10. Brenardez, M., Pastoriza, L., Sampedro, G., Herrera, J. J. and Cabo, M. L. Modified Methods for the Analysis of Free Fatty Acids in Fish. J. Agric. Food Chem., 53, 1903-1906 (2003).

Il. Canham, 1. S. and Pacey, G. E. Automated Free Fatty Acids Determination Using Flow Injection Analysis Solvent Extractions. J. Am. Oil Chem. Soc., 64, 1004-1007 (1987).

12. Puchades, R, Suescun, A. and Maquieira, A. Determination of Free Fatty Acids in Foods by Flow Injection. J. Sei. Food Agric., 66, 473-478 (1994).

13. Zhi, Z-L, Rios, A. and Valcârcel, M. An automated Flow ReversaI InjectionlLiquid-Liquid Extraction Approach to the Direct Determination of Total Free Fatty Acids in Olive Oils. Anal. Chim. Acta, 318, 187-194 (1996).

14. Linares, P., Luque de castro, M. D. and Valcârcel, M. Direct Automatic Determination of Free Acidity in Oils Flow-Injection Analysis. Anal. Chim. Acta, 255,431-436 (1989).

15. Nouros, P. G., Constantino s, A. G. and Polissiou, M. G. Automated Flow Injection Spectrophotometric Non-Aqueous Titrimetric Determination of the Free Fatty Acid Content of Olive Oil. Anal. Chim. Acta, 351, 291-297 (1997).

16. Mariotti, E. and Mascini, M. Determination of Extra Virgin Olive Oil Acidity by FIA-Titration. Food Chem., 73,235-238 (2001).

17. Walde, P. A Colorimetric Determination ofFatty Acids as a New Assay of Lipase in Reverse Micelles. J. Am. Oil Chem. Soc., 67, 110-115 (1990).

18. Walde, P. and Nastruzzi, C. Application of a New, Simple and Economic Colorimetric Method for the Determination of Non-Esterified Fatty Acids in Vegetable Oils. Food Chem., 39, 249-256 (1991).

• 1, 1

19. Takamura, K. and Hayakaw~: Y. betermination of Traces of Free Acid by the Polarographic Reduction of Quinones. Microchem. J., 17,546-550. (1972).

20. Kusu, F., Fuse, T. and Takamura, K. Voltammetric Determination of Acid Values of Fats and Oils. J. AOAC Int., 77, 1686-1689 (1994).

2l. Fuse, T., Kusu, F. and Takamura, K. Voltammetric Determination of Free Fatty Acid Content in Fats and Oils. Bunseki Kagaku, 44, 29-33 (1995).

22. Fuse, T., Kusu, F. and Takamura, K. Determination of Higher Fatty acids in ails by high-Performance liquid Chromatography with Electrochemical detection. J. Chromatogr., 764, 177-182 (1997).

23. Fuse, T., Kusu, F. and Takamura, K. Determination of Acid Values in Fats and ails by Flow Injection Analysis with Electrochemical Detection. J. Pharm. Biomed. Anal., 15, 1515-1519 (1997).

32

Page 50: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

~--

24. Takamura, K., Fuse, T., Arai, K. and Kusu, F. A Review ofa New Voltammetric Method for Determining Acids. J. Electroanal. Chem., 468,53-63 (1999).

25. Kotani, A., Kusu, F. and Takamura, K. New Elecrochemical Detection Method in High-performance Liquid Chromatography for Determining Free Fatty Acids. Anal. Chim. Acta, 465, 199-206 (2002).

26. Lapshina, T. M., Tur'yan, Y. I. and Danilchuk, S. I. pH-Metric Determination of Acid Numbers for Oils. J. Anal. Chem. USSR, 46, 1150-1158 (1991).

27. Tur'yan, Ya. I., Berezin, O. Yu., Kuselman, I. and Shenhar, A. pH-Metric Determination of Acid Values in Vegetable Oîls without Titration. J. Am. Oil Chem. Soc., 73, 295-301 (1996).

28. Tur'yan, Ya. I., Strochkova, E., Berezin, O. Yu., Kuselman, I. and Shenhar, A. pH-Metric Determination of Acid Numbers in Petroleum without Titration. Talanta, 47,53-58 (1998).

29. Kuselman, I,.Tur'yan, Ya. I., Burenko, I. and Anisimov, B. pH-Metric Determination of Acid Values in Oilseeds without Titration. Talanta, 49, 629-637 (1999).

30.

31.

32.

33.

34 ..

35.

36.

37.

38.

39.

40.

Strochkova, E. Tur'yan, Ya. I. and Kuselman, I. pH-Metric Determination for Acid Number in Hydraulic Oils without Titration. Talanta, 50, 1135-1139 (1999).

Lingane, J. J. Automatic Potentiometric Titrations. Anal. Chem., 20, 285-292 (1948).

Muller, R H. Electronic Trigger Circuit for Automatic Potentiometric and Photometrie Titrations. Anal. Chem., 20, 795-797 (1948).

Shaffer, P. A., Briglio, A. and Brockman, A. J. Instrument for Automatic Continuous Titration. Anal. Chem., 20, 1008-1014 (1948).

Delahay, P. Electrical DifferentiaI Method of Measurement in Electrotitrations: Application to Potentiometric Titrations. Anal. Chem., 20, 1212-1215 (1948).

Furman, N. H. Potentiometric Titrations. Anal. Chem., 22, 33-41 (1950).

http://www.mt.comlmtlresourcedetail/appEdStyle.jsp?m=t&key=I2Mjg4NjM1Mj

Vaughan, G. A. and Swithenbank, J. 1., Acetone as Solvent in the Enthalpimetric Titration of Acidic Substances. Analyst, 90, 594-599 (1965).

Thomas, K. S. Analysis of FF A in Edible Oils by Catalyzed End-Point Thermometric Titrimetry (CETT). J. Am. Oil Chem. Soc., 80,21-24 (2003).

Lanser, A. c., List, G. R, Holloway, R K. and Mounts, T. L. FTIR Estimation of Free Fatty Acid Content in Crude Oils Extracted from Damaged Soybeans, J. Am. Oil Chem. Soc., 68, 448-449 (1991).

Ismail, A. A., van de Voort, F. R and Sedman, J. Rapid Quantitative Determination of Free Fatty Acids in Fats and Oils by FTIR Spectroscopy. J. Am. Oil Chem. Soc., 70, 335-341 (1993).

33

Page 51: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

41. Man, Y. B., Moh, M. H. and van de Voort, F. R Determination of Free Fatty Acids in Crude Palm Oil and Refined-Bleached-Deodorized Palm Olein Using Fourier Transform Infrared Spectroscopy, J. Am. ail Chem. Soc., 76, 485-490 (1999).

42. Bertran, E., Blanco, M., Coello, J., Iturriaga, H., Maspoch, S. and Montoliu, 1. Determination of Olive Oil Free Fatty Acid by Fourier Transform Infrared Spectroscopy, J. Am. ail Chem. Soc., 76, 611-616 (1999).

43. Verleyen, T., Verhe, R, Cano, A, Huyghebaert, A and De Greyt, W. Influence of Triacylglycerol Characteristics on the Determination of Free Fatty Acids in Vegetable Oils by Fourier Transform Infrared Spectroscopy, J. Am. ail Chem. Soc., 78,981-984 (2001).

44. Caiiada, M. J. A., Medina, A R and Lendl, B. Determination of Free Fatty Acids in Edible Oils by Continuous-Flow Analysis with FT-IR Spectroscopic Detection, Appl. Spectrosc., 55, 356-360 (2001).

45. Zhang, H-Z and Lee, T-C. Rapid Near-Infrared Spectroscopic Method for the Determination of Fatty Acid in Fish and Its Application in Fish Quality Assessment. J. Agric. Food Chem., 45, 3515-3521 (1997).

46. Cozzolino, D., Murray, 1., Chree, A. and Scaife, J. R Multivariate Determination of Free Fatty Acids and Moisture in Fish Oils by Partial Least-Ssquares Regression and Near-Infrared Spectroscopy. LWT Food Sei. Technol., 38, 821-828 (2005).

47. Man, Y. B. and Moh, M. H. Determination of Free Fatty Acids in Palm Oil by Near-Infrared Reflectance Spectroscopy. J. Am. ail Chem. Soc., 75, 557-562 (1998).

48. Muik, B., Lendl, B., M-D, A and Caiiada, M. J. Direct, Reagent-Free Determination of Free Fatty Acid Content in Olive Oil and Olives by Fourier Transform Raman Spectrometry. Anal. Chim. Acta, 487, 211-220 (2003).

49. Mailer, R J.' Rapid Evalua~ion, of Olive Oil Quality by NIR Reflectance Spectroscopy. J. Am. Oil Chem. 'Soc., 81, 823-827 (2004).

50. Man, Y. B. C and Mirghani, M. E. S., Rapid Method for Determining Moisture Content in Crude PaIm Oil by Fourier Transform Infrared Spectroscopy. J. Am. ail Chem. Soc., 77, 631-637 (2000).

51. Official Methods and Recommended Practices of the American ail Chemists' Society (AOCS, Champaign, IL, 1989), 4th edn.

52. Scholz, E. Karl Fischer Titrations of Aldehydes and Ketones. Anal. Chem., 57, 2965-2971 (1985).

53. Fischer, F. and Schiene, R Dead-Stop Determination of Traces of Water in Aliphatic Ketones with Karl Fischer Solution. Zeitschrift fuer Chemie, 4 (2), 69-70 (1964).

34

Page 52: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

54. Klimova, V. A., Shennan, F. B. and L'vov, A. M. Micro-Detennination ofWater by Fischer Reagent Prepared with N,N'-Dimethylfonnamide Instead of Methanol. Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, 12, 2599-2602 (1967).

55. Mitsubishi Chemical Industries, Gennan Patent Application DE 30 40 474 (1980).

56. Hydranal Composite 5K, leaflet from Riedel de Haen AG, D-3016 seeize.

57. Nordin-Andersson, 1. and Cedergren, A. Coulometric Detennination of Trace Water in Active Carbonyl Compounds Using Modified Karl Fischer Reagents. Anal. Chem., 59, 749-753 (1987).

58. Bizot, J. Automatic Method for Coulometric Detennination of Minute Quantities ofWater. Bull. Soc. Chim. France, 151- 157 (1967).

59. Cedergren, A. Reaction Rates Between Water and some Modified Rapidly­Reacting Karl Fischer Reagents. Talanta, 25,229-232 (1978).

60. Scholz, E. Karl-Fischer reagents without pyridine. 3. Accuracy of the Detennination ofWatèr. Fresenius Z. Anal. Chem., 306, 394-396 (1981).

61. Oradd, C. and Cedergren, A. Reaction Rates of Water in Imidazole-Buffered Methanolic Karl Fischer Reagents. Anal. Chem., 66, 2603-2607 (1994).

62. Verhoef, J. C. and Bahrendrecht, E. Mechanism and Reaction Rate of the Karl­Fischer Titration Reaction. Part 1. Potentiometric Measurements J. Electroanal. Chem., 71, 305-315 (1976).

63. Cedergren, A. and Oradd, C. Coulometric Study of the Influence of Aldehydes and Ketones on the Karl Fischer Titration of Water Using Standard Pyridine !Methanol Reagents. Anal. Chem., 66,2010-2016 (1994).

64. Oradd, C. and Cedergren, A. Coulometric Study of the Recovery Rates for Karl Fischer Titration of Water in Aldehydes and Ketones Using Rapidly Reacting Methanolic and 2-Methoxyethanolic Reagents. Anal. Chem., 67, 999-1004 (1995).

65. Fischer, K. A New Methoçl forth'e'Analytical Detennination of the Water Content of Liquids and Solids: Angew. Chem., 48, 394-396 (1935).

66. Cedergren, A. Coulometric Study of Reaction Rates of Water in Pyridine and Imidazole-Buffered Methanolic Karl Fischer Reagents Containing Chlorofonn. Anal. Chem., 69, 3682-5687 (1996).

67. Isengard, H-D and Kerwin, K. ProposaI of a New Reference Method for Detennining Water Content in Butter Oil. Food Chem., 82, 117-119 (2003).

68. Margolis, S. Amperometric Measurement of Moisture in Transfonner Oil using Karl Fischer Reagents. Anal. Chem., 67, 4239-4246 (1995).

69. Katoh, H. Fujimoto, Y. and Kuwata, Y. A Study of the Coulometric Cathode Reaction in the Karl Fischer Reaction. Anal. Sei. 7,299-302 (1991).

70. Scholz, E. Karl Fischer Coulometry - the Cathode Reaction. Fresenius J. Anal. Chem. 348,269-271 (1994).

35

Page 53: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

71. Cedergren, A. and Jonsson, S. Diaphragm-Free Cell for Trace Determination of Water Based on the Karl Fischer Reaction using Continuous Coulometric Titration. Anal. Chem. 69,3100-3108 (1997).

72. Cedergren, A. and Jonsson, S. Progress in Karl Fischer Coulometry using Diaphragm-Free Cells. Anal. Chem. 73,5611-5615 (2001).

73. Dahms, H. High Current Cou1ometric KF Titrator. US Patent 5300207 (1994).

74. Nichugovskii, G. F. Cou1ometric Cell ofSimplified Design for Titration ofWater in a Fischer's Reagent Medium Zh. Anal. Khim. 32,2124-2126 (1977).

75. Katoh, H. Fujimoto, Y. and Kakuda, M. Effects of the Cathode Current Density in the Karl Fischer Coulometric Titration Method. Anal. Sei. 8,575-577 (1992).

76. Nichugovskii, G. F. and Yakolev, P. P. Coulometric Determination ofWater with Fischer's Reagent in a One-Chamber Cell. Zh. Anal. Khim., 33, 684-687 (1978).

77. Nordmark, U. and Cedergren, A. Conditions for Accurate Karl Fischer Cou1ometry using Diaphragm-Free Cells. Anal. Chem. 72, 172-179 (2000).

78. Nordmark, U., Rosvall, M. and Cedergren, A. Optimum Conditions for Pulse Generation in Daphram-Free Karl Fischer Coulometry. Fresenius J. Anal. Chem., 368, 456-460 (2000).

79. Nordmark, U. and Cedergren, A. Progress in Pulsed-Current Karl Fischer Coulometry using Diaphragm-Free Cells. Fresenius J. Anal. Chem., 367,519-524 (2000).

80. Borsten, H. and van Lamoen, F. L. J. Potentiometric Method for Karl Fischer Titration. Anal. Chem., 27, 1639-1640 (1955).

81. Brown, J. F. and Volume, W. F. An Automatic Fischer Titration Unit for Laboratory Use. Analyst, 81, 308-315 (1956).

82. Cedergren, A. Cou1ometric Trace Determination of Water by using Karl Fischer Regent and Potentiometric End-Point Detection. Talanta, 21, 367-375 (1974).

• j • I~ , 1

83. Lindbeck, M. R.and Freund,' H: Microdetermination of Water, using Rapid Controlled Potential Cou1ometry in a Karl Fischer System. Anal. Chem., 37, 1647-1650 (1965).

84. Cedergren, A. Comparison between Amperometric and True Potentiometric End­Point Detection in the Determination of Water by the Karl Fischer Method. Talanta, 21, 553-563 (1972).

85. Harris, D. C. Quantitative Chemical Analysis, 5th edn., W. H. Freeman and Company, New York, 1999, pp. 498-499

86. Cedergren, A. Comparison between Bipotentiometric and True Potentiometric End-Point Detection using Rapidly Reacting Karl Fischer Reagents. Anal. Chem., 68,3679-3681 (1996).

87. Rosvall, M., Landmarl, L. and Cedergren, A. Computer-Controlled, Coulometric Karl Fischer Systems for Continuous Titration of Water Based on Zero Current Potentiometry. Anal. Chem., 70, 5332-5338 (1998).

36

Page 54: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

88. Bakker, E. and Pretsch, E. Potentiometry at Trace Levels. Trends Anal. Chem., 20, 11-19 (2001).

89. Connors, K. A. and Higuchi, T. Spectrophotometric Detection of the End Point in the Karl Fischer Titration ofWater. Chemist-Analyst, 48, 91-92 (1959).

90. Dahms, H. Method and apparatus for colorimetrie analysis. US Patent 4005983 (1977).

91. Kâgevall, 1., Âstrom, O. and Cedergren, A. Determination of Water by Flow­Injection Analysis with Karl Fischer Reagent. Anal. Chim. Acta, 114, 199-208 (1980).

92. Dahms, Harald. Method for colorimetricdetermination of water using Karl Fischer reagents and optical density measurements at two different wavelengths. US Patent 4786602 (1988).

93. Dahms, Harald. Improved Colorimetrie Karl Fischer Water Determination. PCT Int. Appl., WO 9221969 (1992).

94. Dahms, Harald. Sealed Vials Containing Improved Karl Fischer Solutions and Process for Water Determination using these Vials. US Patent 5179024 (1993).

95. Kâgevall, 1., Âstrom, O. and Cedergren, A. Minimization of Interference Effects from Iodine-Consuming Samples in the Determination of Water with the Karl Fischer Reagent in a Flow-Injection System. Anal. Chim. Acta, 132, 215-218 (1981).

96. Escott, R. E. and Taylor, A. F. Determination of Water by Flow Injection Analysis using Karl Fischer Reagent with Electrochemical Detection. Analyst, 110, 847- 849 (1985).

97. Dantan, N., Kroning, S., Frenzel, W. and Küppers, S. Comparison of Spectrophotometric and Potentiometric Detection for the Determination of Water using Karl Fischer Method under Flow Injection Analysis Conditions. Anal. Chim. Acta, 420, 133-142 (2000).

• , ,,' i il .1

98. Margolis, S. Effect ofHydrocarbon Composition on the Measurement ofWater in Oils by Coulometric and Volumetrie Karl Fischer Methods. Anal. Chem., 70, 4264-4270 (1998).

99. Larsson, W. and Cedergren, A. Coulometric Karl Fischer Titration of Trace Water in Diaphragm-Free Cells. Talanta, 65, 1349-1354 (2005).

100. Smith, B. Infrared Spectral Interpretation: A Systematic Approach, CRC Press, Boca Raton, FL, 1999, Chap. 1, p. 22.

101. van de Voort, F. R., Ismail, A. A., Sedman, J., Dubois, J. and Nicodemo, T. The Determination of Peroxide Value by Fourier Transform Infrared Spectroscopy. J. Am. Oil Chem. Soc., 71, 815-826 (1993).

102. Ma, K., van De Voort, F. R., Sedman, J. and Ismail, A. A. Stoichiometric Determination of Hydroperoxides in Fats and Oils by Fourier Transform Infrared Spectroscopy. J. Am. Oil Chem. Soc., 74, 897-906 (1997).

37

Page 55: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

103. van de Voort, F. R., Sedman, J., Yaylayan, V., Saint Laurent, C. and Mucciardi, C. Quantitative Detennination of Moisture in Lubricants by Fourier Transfonn Infrared Spectroscopy. Appl. Spectrosc., 58, 193-198 (2004).

104. Vomheder, P. F. and Brabbs, W. J. Moisture Determination by Near-Infrared Spectrometry. Anal. Chem., 42, 1454-1456 (1970).

105. Keyworth, D. A. Detennination of Water by Near-Infrared Spectrophotometry. Talanta, 8, 461-469 (1961).

106. Rader, B. R. Determination of Moisture in Dried Vegetables and Spices. J. Assoc. Offic. Anal. Chem., 49, 726-730 (1966).

107. Gold, H. J. General Application of Near-infrared Moisture Analysis to Fruit and Vegetable Materials. Food Technol., 18(4), 184-185 (1964).

38

Page 56: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

CHAPTER3

NEW FTIR METHOD FOR THE DETERMINATION OF FREE FATTY ACID IN

OILS

3.1. ABSTRACT

A rapid, practical and accurate FTIR method for the determination of free fatty

acids (FF A) in edible oils has been developed. Analogous to the AOCS titration

procedure, FTIR FF A determination is effected by an acid/base reaction but directly

measures the product formed rather than utilizing an endpoint based on an electrode

potential or color change. A suspension of a weak base, potassium phthalimide (K­

phthal) in I-PrOH, is used to convert FFA present in oils to their carboxylate salts

without causing oil saponification, and differential spectroscopy is used to circumvent

matrix effects. Samples are tirst diluted with I-PrOH, then split, with one half treated

with the K-phthal reagent and the other half with I-PrOH (blank reagent), their spectra

collected, and a differential spectrum obtained to ratio out the invariant spectral

contributions from the oil sample. Quantitation of the percentage of FF A in the oil,

expressed as % oleic acid, based on measurement of the peak height of the v (COO")

absorption of the FF A salt formed yielded a calibration with a standard error of <0.020 %

FF A over the range of 0-4%. The method was validated by standard addition and the

analysis of Smalley check samples, the results indicating that the analytical performance

of the FTIR procedure is as good as or better than that of the standard titrimetric . "

procedure. As structured, the FTIR procedure is a primary method, as calibration is not

dependent on reference values provided by another method, and has performance criteria

which could lead to its consideration as an instrumental AOCS procedure for FF A

determination. The FTIR portion of the analysis is automatable, and a system capable of

analyzing ~60 samples/h has been developed that could be of benetit to laboratories that

carry out a large number of FF A analyses per day.

3.2. INTRODUCTION

FTIR spectroscopy is playing an increasingly important role in the analysis of

edible oils by providing simpler and more rapid techniques for determining common oil

quality parameters (1). FFA are common triacylglycerol (TAG) hydrolysis products in

39

Page 57: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

crude oils and are fonned to sorne extent in refined oils as a result of oxidation or TAG

degradation during frying, impairing oil quality and functionality. Chemically, FFAs are

less stable than TAG and therefore more likely to oxidize and cause rancidity (2). The

standard method commonly used for FF A analysis is based on the titration of an oil

dissolved in alcohol with a strong base to a phenolphthalein endpoint (3,4). Although

simple, titrimetric methods are tedious, consume substantial amounts of solvent, and can

be problematic when dark crude oils are analyzed. The first application of FTIR

spectroscopy for FFA analysis was reported by Lanser et al. (5) in 1991. In this method,

which was developed for crude soybean oil, the FF A content was estimated from the FF A

v (C=O) band at 1710 cm- l. Owing to the overlap ofthis band with the very strong TAG

ester carbonyl absorption at 1746 cm-l, spectral deconvolution over the 2000-1600 cm- l

range was used to mathematically enhance spectral resolution. A calibration was derived

by spiking oleic acid into soybean oil at levels of 0.1-5% and yielded predictions of the

FF A content of soybean oils that matched the values obtained using the AOCS titrimetric

method to within ±0.5 percentage points. However, because the FTIR spectra were

acquired by simply placing each sample between two KBr windows, without the use of

an internaI standard, the accuracy of this method was limited by the resulting variability

in pathlength. In 1993, Ismail et al. (6) investigated two different FTIR approaches to the

quantitative detennination of FF A in edible oils. The first was based on measuring the

carboxylic acid v (C=O) band at 1711 cm-l in spectra acquired from oil samples applied

in their neat fonn ont9 an attenuated totaJ reflectance (ATR) crystal. Both calibration and , .

sample spectra were ratioed against the' ~pectrum of an FFA-free oi! of the same type as

the oi! being analyzed to reduce matrix effects. The second approach was an indirect

method based on the use of KOHlmethanol to extract the FF A present in the oi! and

convert them to their salts, followed by measurement of the v (COO") absorption band at

1570 cm- l in the spectrum of the methanol phase. This indirect method enhanced the

sensitivity of the analysis by concentrating the FF A in a small volume of methanol and

by utilizing an absorption band that is in a region free from major oïl spectral

Interferences. However, it had the disadvantage of requiring an additional procedural

step, and sorne saponification of oils by the KOHlmethanol reagent added was noted,

resulting in overestimation of the FF A content of the original sample. Verleyen et al. (7)

40

Page 58: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

devised a more rigorous version of the method of Lanser et al. using peak height

measurements at 1711 cm-1 to develop workable calibrations for a variety of oils but

ultimately concluded that the calibrations were strongly oil dependent. Two other

publications have dealt with FF A analysis by FTIR spectroscopy, specifically in relation

to palm and olive oil, respectively (8,9), both using partialleast-squares regression (PLS)

to develop relationships between spectral changes and results obtained by standard

methods. The most sophisticated methodology is that of Cafiada et al. (10), who

developed an automated FTIR-based continuous-flow analysis system capable of

analyzing ~40 samples per hour using the indirect approach described by Ismail et al. (6).

However, even within the ~90 s required for analysis in this automated system, sorne

saponification of the oil can occur. In general, the main drawback associated with the

direct FTIR methods is their oil dependency, while the indirect FTIR methods are limited

by the possibility of errors due to saponification caused by the KOHImethanol reagent.

Hence, a simple, reliable, and robust FTIR method for FF A analysis is stilliacking.

The McGill IR Group recently developed an FTIR-based instrumental method to

replace the American Society for Testing and Materials (ASTM) titrimetric procedures

for the determination of acid number (AN) in mineraI and ester-based lubricating oils

(11). These ASTM methods are similar to those traditionally employed for the

determination of FF A content in edible oils except that they measure not only carboxylic

acids but also a variety of other acidic constituents, organic or inorganic, that accumulate

in lubricating oils either as a result, of Qxidation or as combustion by-products (12). " • 1

Because the FTIR AN method was specifically designed to meet the requirements of

lubricant analysis, it is not directly applicable to edible oils owing to their different

spectral characteristics. However, elements of this methodology have been adapted to

develop a new method for FF A determination that overcomes the limitations of both the

direct and indirect FTIR methods previously developed for FF A analysis. This paper

describes the princip les of this method as well as their practical implementation and

provides an evaluation of its performance by standard addition as weIl as by employing

AOCS SmaIley check samples.

41

Page 59: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

3.3. EXPERIMENTAL PROCEDURE

Reagents and Standard Methods. AlI reagents used were of analytical grade.

Potassium phthalimide (K-phthal, 99+%) and hexanoic acid (99%) were obtained from

Aldrich (St. Louis, MO); I-propanol (l-PrOH) and iso-propanol were purchased from

Fisher Scientific Ltd. (Nepean, ON, Canada). AlI edible oils were obtained localIy and

samples were analyzed for FFA using AOCS method Ca 5a-40 (3). Mineral oil (C-171

polyalphaolefin) was obtained from Thermal-Lube (Montreal, QC, Canada) and used for

reagent blank determinations. A series of five oils pre-analyzed for FF A content were

obtained ,from the AOCS Smalley Check Sample program.

Instrumentation. The instrument used for this study was a Bornem WorklR

spectrometer (Bornem, Quebec, QC, Canada) equipped with a DTGS detector and purged

with dry air from a Balston dryer (Balston, Lexington, MA). Samples were analyzed by

aspirating them into a 500 flm Caf 2 transmission flow cell mounted on a sample shuttle

(Dwight Analytical, Toronto, ON, Canada). The spectrometer was controlled by an IBM­

compatible Pentium 150-MHz PC running under proprietary Windows-based UMPlRE®

(Univers al Method Platform for InfraRed Evaluation) software (Thermal-Lube, Pointe­

Claire, QC, Canada). AIl spectra were collected by co-adding 16 scans at a resolution of 8

cm-1 and a gain of 1.0.

Preparation of Calibration Standards. A series of eight standards covering the

range 0-4% FF A (expressed as perce~t.~g~ of oleic acid) were prepared by gravimetric 1 i f

addition of hexanoic acid to '3. retlned and deodorized soybean oil. The calibration curve

was obtained by linear regression of % oleic acid (%FF A) against the peak heights in the

FTIR spectra recorded for the standards by following the sample preparation and

analytical protocols described below.

Sam pIe Preparation for FTIR Analysis. Six grams of the oil sample was mixed

with 3 ml of I-PrOH in a 20-mL vial. Three-milli1iter aliquots of the diluted oïl were

placed in two centrifuge tubes, labeled BR (blank reagent) and RR (reactive reagent), to

which were added 7 mL of I-PrOH and 7 mL ofK-phtha1l1-PrOH (20 gIL), respectively.

AIl tubes were capped, shaken on a vortex mixer, and then centrifuged for a minimum of

5 min at 6000 rpm in a clinical centrifuge. It should be noted that K-phthal is virtually

42

Page 60: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

insoluble in I-PrOH and was dispensed via bottle re-pipette ~s a fine dispersion which

was maintained by continuous and vigorous agitation on a magnetic stirrer.

Analytical Protocol. The transmission flow cell was loaded with ~2 mL of the BR

sample and its single-beam spectrum was recorded. After the cell was flushed with iso­

propanol, -2 mL of the RR sample was loaded into the cell and its single-beam spectrum

was recorded and ratioed against that of its corresponding BR sample to pro duce a

differential spectrum. For quantitation of FFA, the peak height of the carboxylate v

(COO") band at 1570 cm-l in the differential spectrum was measured relative to a baseline

point at 2150 cm- l. A schematic diagram of the sample preparation procedure and

analysis is illustrated in Figure 3.1.

Diferential Spectrum

Figure 3.1. Schematic diagram illustrating the sample preparation procedure and the steps in the analytical protocolleading to the differential spectrum.

43

Page 61: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

Validation. The FTIR method was vaHdated by standard addition of FF A to

soybean and corn oil and by analyzing AOCS Smalley check samples, which inc1uded

five types of oils (crude coconut, crude corn, crude safflower, cottonseed, and marine

oils). For the standard addition experiments, an FFA mixture was prepared by

saponifying olive oil with 50% w/v KOH followed by titration with 6M HCI to

regenerate the FF A (13), extraction into n-hexane, removal of the solvent using a rotary

evaporator, and titration of the residue to determine %FF A (expressed as % oleic acid).

Soybean and corn oils were then spiked (w/w) with six levels of this FF A mixture. These

samples and the Smalley check samples were analyzed for their FF A content by the FTIR

method as well as by the AOCS titrimetric method. Reproducibility was evaluated as the

standard deviation around the mean oftriplicate analyses (SDr). For the standard addition

experiments, accuracy was àssessed in terms of mean difference (MDa) and standard

deviation of the differences (SDDa) with respect to the gravimetrically spiked amounts of

FF A. In the case of the Smalley check samples, the FTIR results were compared to the

results of certified laboratories that had analyzed the samples by the AOCS titrimetric

procedure, using the statistical data (mean, min and max) provided with the samples.

3.4. RESULTS AND DISCUSSION

3.4.1. Analytical Concepts

As noted, FTIR methods developed to date for FF A analysis have been either

matrix dependent or l?rone to saponifica~ion errors. In FTIR methodology developed for

AN analysis in lubricants, the weak baSe K-phthal, a dicarbonyl compound, served as a

signal-transducing reagent by reacting with all organic and inorganic acids present to

form a single product, phthalimide, allowing AN to be determined from the intensity of

phthalimide's strong v (C=O) band at 1729 cm-l, with differential spectroscopy then

being used to eliminate the oil matrix effects (12), This method is not directly applicable

to edible oils, because the 1729 cm-l phthalimide band is masked by overwhelming

absorptions of the ester linkages of the TAG, whereas the second phthalimide v (C=O)

band at 1773 cm-l is too weak to measure FFA concentrations of <1 %. However, for the

analysis of FF A, signal transduction is not required per se, only the stoichiometric

conversion of FF A to their salts, which can then be quantitated directly by measurement

44

Page 62: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

of their v (COO-) absorption. Moreover, the base used should be too weak to hydrolyze

TAG, unlike the KOH employed in previous work (6,10). K-phthal (PKa = 9.9) was found

to meet this criterion, with the additional advantage that I-PrOH could be used to deliver

this reagent into oils, thereby eliminating phase separation, with the FF A salts remaining

soluble in the I-PrOHioil mixture. The stoichiometric reaction of K-phthal with FF A is

presented below:

RCOO H :::::;;~r==:::!"'~

o

e® NH + RCOO K

o

As in the lubricant AN methodology, the problem of matrix effects, which may

arise in analyzing different oil types or result from the presence of various minor

constituents in the samples analyzed, is addressed by utilizing differential spectroscopy.

This approach involves splitting the sample into two parts; one part is then treated with

K-phthal in I-PrOH (designated RR) while the other portion is treated only with an

equivalent amount of I-PrOH (designated BR) and serves as a reference for the reacted

sample. Since the spectral features of the oil are invariant in the spectra of the BR- and

RR-treated portions, 'they are cancelled out in the differential spectrum and only the

spectral changes associated with the aeid/base reaction are left for evaluation (Figure

3.2).

3.4.2. Calibration and Stability of the Reaction

Figure 3.2 illustrates typical differential spectra obtained when calibration

standards (soybean oïl spiked with hexanoic aeid over a range of 0-4% FFA) were

analyzed using the protocol described above. As the amount ofhexanoic acid increases in

each subsequent standard, the v (COO) signal at 1570 cm-1 and that of the phthalimide v

(C=O) band at 1773 cm-1 rise concurrently. The loss of the v (C=O) band of hexanoic

acid, which would manifest itself as a strong negative band around 1710 cm-l, is largely

45

Page 63: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

:/"'--', lost in the noise (1770-1700 cm- l) resulting from the subtraction of the off-scale bands of

the ester linkage of the oil.

1570 cm-1

1.0 1773 cm- l

0.8

~ y 0.6 = = ~ .. 0 Vi 0.4 ..c <

0.2

0.0~~r11

1800 1750 1700 1650 1600 1550

Wavenumber (cm-1)

/ i f

Figure 3.2. Potassium phthalimide as a reagent to carry out the acid/base reaction. Spectra were recorded in a 500 Ilm cell at 8 cm-l resolution. The bands at 1773 and 1570 cm- l are due to the phthalimide and hexanoate, respectively, formed in the reaction. The region between 1760 and 1710 cm- l is obscured by noise in the differential spectrum because of the intense oïl absorption in this region.

A mean plot with SD bars for three sets of hexanoic acid calibration standards

obtained by measuring the v (COO-) band at 1570 cm-l referenced to 2150 cm-1 vs. FFA

concentration expressed as % oleic acid is presented in Figure 3.3. The overalllinear

regression equation obtained for the composite calibration was:

46

Page 64: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

% FF A(OIeic) = 4.281 * A(1570/2150) - 0.0436

R2 = 0.999 SD = 0.020 [3.1 ]

Figure 3.3. Composite calibration curve obtained from the differential spectra of three independent sets of hexanoic acid-spiked soybean oil standards of the type illustrated in Figure 3.1.

The linearity of the composite calibration plot, with an intercept well within 3 x

the regression SD, and the overall SD of -0.02% FF A indicate that the three individual

calibrations that were performed gave highly consistent results. However, it was noted

that calibrations tended to drift over a period of a week, and this was found to be due to

changes in the K-phthal reagent, making it necessary to perform a reagent blank

correction. Although K-phthal is practically insoluble in I-PrOH and is delivered as a

suspension, small amounts do solubilize slowly over time. Because the solubilized K­

phthal has absorptions that overlap with the v (COO) band used to quantitate the FFA

salts, it can introduce a measurable bias into FF A measurements over time, albeit taking

several weeks to develop in a freshly prepared reagent. To account for this reagent

47

Page 65: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

background signal, a mineraI oil, which does not contain any FF A, is run as a blank to

compensate for absorptions contributed by any solubilized K-phthal. The apparent FFA

contribution of the blank is subtracted from the values obtained for all oil samples

subsequently analyzed to account for any changes in K-phthal concentration.

Duplicate analyses of calibration standards were conducted 24 h apart to confirm

that K-phthal does not attack TAG. Based on the MDr and SDDr of 017% and 0.024%,

respectively, it was conc1uded that no significant saponification took place over that time

period. This finding was corroborated by similar studies with samples of various refined

oils.

3.4.3. Evaluation and Validation

To establish the efficacy of the methodology developed, standard addition

experiments were carried out by adding a pre-prepared mixture of FF A obtained from

olive oil to both refined corn and soybean oils, which were then analyzed in triplicate by

the FTIR and the titrimetric AOCS method. The data obtained using these two methods

are summarized in Table 3.1, with the accuracy ofboth methods being assessed by using

the gravimetrically added amounts ofFF A as the reference values.

In terms of overall reproducibility, the FTIR method had a mean SDr of 0.029%,

slightly better than that of the AOCS procedure (0.038%). In terms of overall accuracy

relative to gravimetric standard addition, both methods have small positive biases,

slightly greater than. the overall SDr, indicating that both the soybean and corn oils , " " ,

contained traces ofFFA prior to standard addition. For each method, the value of SDDa,

which is a measure of the variability around the MDa, is of the same order of magnitude

as the SDr value, with the FTIR method again performing slightly better. Figure 3.4

present a composite plot of all the FF A results obtained for both oils by both methods

against the reference gravimetric data. The regression equation for the composite data

illustrated in Figure 3.4 is:

FFA(IRITitration) = 0.99996 FFA(StandardAddition) + 0.047

R2 = 0.999 SD =0.030 [3.2]

48

Page 66: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

Table 3.1. Results of Triplicate Analyses ofOils Spiked with Various Amounts an FF A Mixture by the AOCS Titrimetric and FTIR Methodsa

FFA spiked AOCSMethod FTIRMethod Oil

(%W/W)b Mean SD Mean SD

0.000 0.028 0.003 0.015 0.032

0.501 0.524 0.015 0.526 0.019

Soybean 1.002 1.063 0.034 1.042 0.035

1.508 1.543 0.061 1.549 0.031

2.003 2.033 0.029 2.056 0.032

2.490 2.531 0.055 2.505 0.037

0.000 0.057 0.011 0.066 0.020

0.500 0.546 0.009 0.573 0.027

Corn 0.999 1.120 0.053 1.038 0.023

1.475 1.573 0.068 1.512 0.005

2.002 2.094 0.076 1.991 0.088

2.502 2.578 0.051 2.548 0.002

Mean SDr 0.038 0.029

MDa 0.059 0.036

SDDa 0.030 0.022

"Abbreviations: SDr. SD for reproducibility; MD •• mean difference for accuracy; SDD •• SD of the differences for accuracy.

bExpressed as % oleic acid.

The plot and regression equation illustrate the excellent concurrence between the

titrimetric and FTIR methods as well as their comparable ability to track the amounts of

FF A added gravimetrically to the two oils.

49

Page 67: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

3.0 • AOCS Soybean oil • AOCS Comoil -.. 2.5 ! 6. FTIR Soybean oil

~ * FTIRComoil

~ 2.0 = --"CI ~ Jo.. 1.5 = c:I.l ~ ~

~ 1.0 -< ~ ~ 0.5

0.0 • 0.0 0.5 1.0 1.5 2.0 2.5

FFA Added (% w/w)

Figure 3.4. Graphical comparison of results obtained by the AOCS and FTIR methods for the oils that were subjected to standard addition of FF A mixture, relative to the gravimetrical data.

To further validate the performance of the FTIR method, five AOCS Smalley

check samples were analyzed in triplicate. The results obtained relative to the analytical

data provided with these oïl samples are presented in Table 3.2. In general, there is , ,

excellent concurrence between the FTIR' mean and Smalley mean FF A values except for

crude coconut oil. Analysis of the Smalley samples in our laboratory by the AOCS

method were also in line with the Smalley means, except again for crude coconut oil,

which produced a value of 0.312 ± 0.005, very much in line with the FTIR result

obtained, suggesting that the FF A content of this sample had changed in the time that had

elapsed since its analysis in laboratories participating in the Smalley check sample

program. Considering the results for the Smalley check samples as weIl as those obtained

by standard addition, it is evident that the FTIR method is capable of producing accurate

and reproducible FF A data independent of oil type and appears to be a valid alternative to

the AOCS titrimetric procedure.

50

Page 68: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

Table 3.2. Results of Triplicate Analysis ofSmalley Check Samples by AOCS Method and FTIR Methoda

AOCS Method FTIRMethod Oil sample

Max Min Mean SD nb Mean SD

Crude safflower oïl 0.500 0.330 0.457 0.036 18 0.467 0.057

Cottonseed oil 0.180 0.040 0.130 0.053 8 0.156 0.058

Crude coconut oil 0.200 0.080 0.136 0.029 20 0.297 0.014

Crude corn oïl 1.720 1.120 1.419 0.124 19 1.431 0.019

Marine oïl 1.970 1.800 1.864 0.044 12 1.810 0.024

aResults expressed as % oleic acid (w/w)

b n = number of laboratories that reported results for each sample.

The FTIR methodology for FF A analysis developed in this work combines the

respective advantages of the direct and indirect approaches previously described in the

literature and overcomes their limitations. As in other indirect approaches, enhanced

sensitivity by comparison with direct measurement of the FF A v (C=O) band is achieved

by reacting the FFA with a base and measuring the v (COO") band of the salt formed.

However, a key difference is the use of a K-phthal suspension in 1-PrOH as a reagent

instead of the KOHImethanol reagent previously used (6,10) as its weakly basic . ,

properties provide a means of avoiding sàponification. Furthermore, the extraction step

previously required in the indirect FF A methods is eliminated because no phase

separation occurs and the FF A salts remain soluble in the I-PrOHIoil mixture. The use of

differential spectroscopy provides a means of minimizing matrix effects by canceling out

the spectral contributions of the oil and contaminants therein, and spectral interferences

arising from the slight solubilization of K-phthal in a non-freshly prepared reagent are

compensated for by performing a reagent blank correction. These combined procedural

elements allow a calibration to be developed by utilizing gravimetrically prepared

standards of hexanoic acid in oil. The net result is that the FTIR method developed is a

primary method, independent of other methods for calibration. It is also rapid and simple

51

Page 69: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

to execute, and the FTIR portion of the analysis has been automated by integrating the

spectrometer with an autosampler (Thermal-Lube Inc., Pointe-Claire, QC, Canada),

allowing for the analysis of up to 60 paired samples per hour. Given that only

measurements at two wavelengths are needed for this analytical procedure, a simple dual­

wavelength filter-based IR instrument could also be used. With these hallmarks of a

sound general method, this new methodology could serve as the basis for an AOCS

instrumental method for the analysis ofFFA in edible oils.

3.5. ACKNOWLEDGMENT

The authors would like to thank Sultan Qaboos University, Oman, for support of

Mr. Alawi and the Natural Sciences and Engineering Research Council (NSERC) of

Canada for financial support of this research. The technical assistance of Ms. Claudia

Mucciardi is also acknowledged.

3.6.

1.

2.

3.

4.

5.

6.

7.

8.

REFERENCES

van de Voort, F. R, Sedman J. and Russin, T. Lipid Analysis by Vibrational Spectroscopy, Eur. J. Lipid Sci. Technol. 103,815-826 (2001).

O'Brien, R D. Fats and Oils: Formulating and Processing for Applications, Technomic Publishing Company, Lancaster, PA, 1998.

Official Methods and Recommended Practices of the American Oil Chemists' Society, 4th edn., 1989, Champaign, AOCS Official Method Ca-5a-40.

Determination of the Free Fatty Acids in European Commission Regulation (EED) No. 2568/91, Official Journal of the European Communities, No. L 248 (5.9.91), p. 6. '

, ' "

Lanser, A C., List, G. R, Holloway, R K., and Mounts, T. L. FTIR Estimation of Free Fatty Acid Content in Crude Oils Extracted from Damaged Soybeans, J Am. Oil Chem. Soc., 68, 448-449 (1991).

Ismail, A. A., van de Voort, F. R and Sedman, J. Rapid Quantitative Determination of Free Fatty Acids in Fats and Oils by FTIR Spectroscopy. J Am. Oil Chem. Soc., 70, 335-341 (1993).

Vedeyen, T., Verhe, R, Cano, A, Huyghebaert, A and De Greyt, W. Influence of Triacylglycerol Characteristics on the Determination of Free Fatty Acids in Vegetable Oils by Fourier Transform Infrared Spectroscopy, J Am. Oil Chem. Soc., 78, 981-984 (2001).

Man, Y. B., Moh, M. H. and van de Voort, F. R Determination of Free Fatty Acids in Crude Palm Oil and Refined-Bleached-Deodorized Palm Olein Using

52

Page 70: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

. ~.

Fourier Transfonn Infrared Spectroscopy, J. Am. ail Chem. Soc., 76, 485-490 (1999).

9. Bertran, E., Blanco, M., Coello, J., Iturriaga, H., Maspoch, S. and Montoliu, 1. Detennination of Olive Oil Free Fatty Acid by Fourier Transfonn Infrared Spectroscopy, J. Am. ail Chem. Soc., 76, 611-616 (1999).

10. Cafiada, M. J. A., Medina, A. R and Lendl, B. Detennination of Free Fatty Acids in Edible Oils by Continuous-Flow Analysis with FT-IR Spectroscopic Detection, App!. Spectrosc., 55, 356-360 (2001).

11. van de Voort, F. R, Sedman, J., Yaylayan, V. and Saint Laurent, C. The Detennination of Acid Number and Base Number in Lubricants by FTIR Spectroscopy. Appl. Spectrosc., 57, 1425-1431 (2003).

12. Kishore N. R A., Guide to ASTM Test Methods for the Analysis of Petroleum Products and Lubricants (ASTM, West Conshohocken, PA, 2000), p. 45~

13. Official Methods of Analysis of the Association of Official Analytical Chemists, 17th edn., Arlington, 2000, Method 972.28 .

53

Page 71: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

...--., l '

BRIDGE

In Chapter 3, an FTIR method for determination ofFFA content in edible oils was

developed by solubilizing oils in I-PrOH containing a weak base and spectrally

evaluating the FF A salts formed, using differential spectroscopy to minimize matrix

effects. The method focused on finding ways to avoid or eliminate problems that were not

overcome by previously developed FTIR methods, such as saponification and calibration

dependency on oil type. Although successful in achieving these goals and providing a

reliable and reproducible instrumental method for the determination of FF A, this new

FTIR method is not significantly better than standard titrimetric procedures in terms of

sensitivity. In Chapter 4, the knowledge and experience gained from this initial work is

used to develop a much more sensitive method which allows for the analysis of low

levels ofFFA in refined edible oils .

54

Page 72: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

CHAPTER4

A NEW METHOD FOR THE ANALYSIS OF LOW LEVELS OF FFA IN

REFINED EDIBLE OILS

4.1. ABSTRACT

This paper summarizes the application of stoichiometric analytical approaches to

quantitative IR analysis and describes the development of a rapid and sensitive Fourier

transform infrared (FTIR) method using such an approach for the determination of low

levels «0.005%) of free fatty acids (FFA) in refined edible oils. The method simply

involves mixing the sample with methanol containing 2 gIL sodium hydrogen cyanamide

(NaHNCN) on a vortex mixer for 30 s to convert the FFA to their salts, centrifuging the

sample to separate the methanol phase containing the FF A salts from the oil, recording

the FTIR spectrum of the upper methanollayer in a 100 Jlm CaF2 transmission flow cell,

and ratioing this spectrum against that of the NaHNCN/methanol solution. The

concentration of FF A salts is determined from the resulting differential spectrum by

measurement of the v (COO) absorbance at 1573 cm-l relative to a reference wavelength

of 1820 cm-l. A calibration spanning the range 0-0.1 % FF A (expressed as oleic acid) was

devised by gravimetric addition of a defined, pure fatty acid to an acid-free oil.

Validation of the method by standard addition of palmitic acid to a variety of oils yielded

an overall standard error of <±0.001% FFA. Comparison oftriplicate FTIR and IUPAC

titrimetric analyses of oils spiked with palmitic acid demonstrated that this FTIR method " , . ' '

was more sensitive, accurate and reproducible than the titration procedure, the latter

having a significant positive bias of ~0.02%. Solventloil consumption in the FTIR

method is 2 mLll 0 g vs. 150 mL/20 g for the titrimetric procedure. The FTIR method

developed is particularly well suited for the determination of the low levels of FF A in

refined oils but can readily be adapted with a simple adjustment of the oi1:methanol ratio

to coyer FFA levels ofup to 4.0%.

4.2. INTRODUCTION

The application of chemometric techniques has represented a major advance in

quantitative IR analysis. In particular, partial-Ieast-squares regression (PLS) has been of

55

Page 73: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

widespread utility (1) as it provides a powerful means of extracting quantitative spectral

information related to component(s) of interest from the spectra of complex samples by

mathematically modeling partially overlapping bands or other sources of spectral

interference. However, as the complexity of the system increases, so does the difficulty of

developing reliable PLS calibration models and we have encountered diverse

circumstances where PLS proved to be unsatisfactory as a calibration technique.

Consequently, an alternative approach to quantitative IR analysis was developed based on

the use of a reagent that reacts stoichiometrically with the component(s) of interest to

pro duce a readily measurable IR signal. Furthermore, by combining this "signal

transduction" approach with differential spectroscopy, quantitation can often be achieved

via a simple univariate calibration.

Table 4.1 illustrates a number of successful applications of this approach. In the

first case, difficulties were encountered in developing a robust PLS calibration model for

determination of the peroxide value (PV) of edible oils. PV is a measure of

hydroperoxides (ROOH) , the primary oxidation products in edible oils, but FTIR

quantitation of these species was complicated by both spectral overlap and hydrogen­

bonding interactions with a large variety of secondary oxidation products and other

components that may be present in edible oils (2). These problems were overcome by

taking advantage of the rapid stoichiometric reaction of hydroperoxides with

triphenylphosphine (TPP) to form triphenylphosphine oxide (TPPO), which has an

isolated and intense absorption band ~at al~owed for the accurate determination of PV " ,

down to <1.0 mequivalents ROOHlkg oil (3).

The next three examples in Table 4.1 concern the analysis of new and used

lubricating oils, which is beset by even more complex matrix effects than those

encountered in oxidized edible oils. In the case of moi sture analysis (4), signal

transduction was carried out by stoichiometrically converting dimethoxypropane through

its reaction with water into acetone, a strong IR absorber, providing an alternative to the

problematic Karl Fischer titration procedure. For the determination of acid number (AN)

and base number (BN) in lubricants, there was the additional complication of the

"structural specificity" of IR analysis as opposed to the "chemical specificity" of the

traditional titrimetric methods employed for these analyses. The latter methods directly

56

Page 74: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

VI -....l

')

Table 4.1. FTIR Analyses Based on the Use ofStoichiometric Reactions for "Signal Transduction"

Analysis· Reagent Stoichiometric Reaction FTIR Measurement

PV (>1.0) Triphenylphosphine ROOH + TPP ~ ROH + TPPO 542 cm-1 [X-sensitive ring-breathing

Edible Oils (TPP) in l-hexanol vibration ofphenyl groups in TPPO]

H20 «1000 ppm) 1,3-Dimethoxypropane 2H20 + DMP ~ CH3COCH3 + 2CH30H + H20 1717 cm-1 [v(C=O) absorption of

Lubricants (DMP) in isooctane - acetone]

AN «0.1 mg KOHlg) Potassium phthalimidè - HA + K+Phthar ~ Ph + A" K+ 1774 or 1727 cm-1 [v(C=O)

Lubricants (K-Phthal) in 1-PrOH' - absorptions of phthalimide]

-

BN «0.1 mg KOHlg) Trifluoroacaetic acid B: + TFA ~ TFA"BH + 1679 cm-1 [v(COO") absorption of

Lubricants (TF A) in 1-PrOH trifluoroacetate]

FFA (>0.2% C 18:1) Potassium phthalimide RCOOH + K+Phthar ~ RCOO- + Ph 1570 cm-1 [v(COO") absorption of

Edible Oils (K-Phthal) in 1-PrOH RCOO-] or 1776 cm-1 [v(C=O)

absorption of phthalimide]

FFA(>0.002% C 18:1) Sodium hydrogen RCOOH + NaHNCN ~ RCOO- + H2NCN 1573 cm} in spectrum ofmethanol

Refined Edible Oils cyanamide in methanol extract [v(COO") absorption of

RCOO1 - --~~

apV = Peroxide Value, H20 = Moisture, AN = Acid Number, BN = Base Number, FFA = Free Fatty Acids.

Rer.

(2)

(3)

(4)

1

(5)

(6)

This work

Page 75: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

measure a large variety of acids and bases, their response being dependent only on the

pKa of the acid or base in relation to the titrimetric endpoint. In contrast, prediction of AN

or BN by direct IR analysis would require the development of calibrations that model all

the species contributing to the acidiclbasic characteristics of lubricating oils, many of

which are undefined. This severe limitation was overcome by reacting all the acidic or all

the basic species with, respectively, a basic or an acidic "signal-transducing" reagent. By

subtracting the spectrum of the unreacted sample from that of the reacted sample, the

spectral changes associated with the acid-base reaction were isolated, and the extent of

conversion of the "signal-transducing" reagent could be directly measured to determine

the AN or BN of the sample (5).

The last two examples cited in Table 4.1 concern the application of similar

concepts to the determination of the free fatty acid (FF A) content of edible oils, the first

of which was described in a previous paper (6). Among the oil quality parameters, FF A

content is a crucial factor associated with the quality and economic value of edible oils,

especially for unrefined high value oils such as olive oil. FF A content is also an important

quality indicator in relation to oil processing and is used to assess deodorizer efficiency

or as an indicator of frying oil quality (7,8). FFA content is most commonly determined

by titration of an oil, dissolved in neutralized ethanol or ethanol/diethyl ether, with a

strong base to a phenolphthalein endpoint (9,10). Although the standardized titrimetric

methods are fairly sensitive, with limits of detection (expressed as percent oleic acid) on

the order of 0.03% being attainable, ,mor~.sensitive methods would be useful for the \' "i f

analysis of refined, bleached and deodorized (RBD) oils, which tend to have FF A levels

of :::;0.05% (7), and could also provide an alternative means of monitoring secondary

oxidation products (11) accumulating in an oil in the form of carboxylic acids. In recent

years, a variety of approaches have been investigated as possible alternatives to the

titrimetric methods employed to determine the FF A content of oils, inc1uding the use of

flow injection systems (12,13) pH metric, potentiometric and colorimetric (14-16)

methods, and chromatographic procedures (17-20) as well as FTIR-based spectroscopic

techniques (21-28). Although many of these offer substantial benefits, in terms of speed

of analysis, amenability to automation, andlor a reduction in the use of solvents and the

attendant environmental problems and disposaI costs, none of them provide a substantive

58

Page 76: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

~'" gain in sensitivity over that attained with titrimetric methods. This paper describes a

simple, robust FTIR method based on the concepts outlined above that is capable of

measuring FF A levels as low as 0.005% in refined oils. Thus, the new methodology

described in this paper exemplifies the means by which the sensitivity of IR analysis can

be substantially enhanced under certain circumstances by the signal

transductionldifferential spectroscopy approach.

4.3. METHODOLOGY

Reagents and Standard Methods. Sodium hydrogen cyanamide (NaHNCN,

99+%), palmitic acid (99%), and anhydrous methanol (MeOH) were obtained from

Aldrich (St. Louis, MO) and were all of analytical grade. Refined edible oils were

purchased locally or obtainedfrom Canamera Foods (Toronto, ON, Canada). The reagent

solution employed in the FTIR FF A analysis was prepared by dissolving NaHNCN in

anhydrous MeOH (2 gIL). This solution was allowed to stand for -4 days or until the v

(C=N) band at 2100 cm-1 completely disappeared before use.

Instrumentation. The FTIR spectrometer used for this study was a Bornem

WorkIR (Bornem, Quebec, PQ, Canada) equipped with a DTGS detector and purged with

dry air using a Balston dryer (Balston, Lexington, MA). The sample-handling accessory

was a valved 100 p,m CaF2 transmission flow cell (Dwight Analytical, Toronto, ON,

Canada). Samples were aspirated into the cell under vacuum, and the cell was flushed

c1ean after each sample with 1 mL of methanol. All spectra were collected by co-adding

32 scans at a resoluti~n of 8"cm- l and' a ~g~in of 1.0. The spectrometer was controlled by

an IBM-compatible Pentium ISO-MHz PC running under proprietary Windows-based

UMPIRE® (Univers al Method Platform for InfraRed Evaluation) software (Thermal­

Lube, Pointe-Claire, PQ, Canada). This software provides programming capabilities so

that repetitive operations can be performed in a specified sequence and designated

spectral data collected and processed through a calibration equation, thereby automating

the analysis to provide direct output ofFFA data.

Preparation of Calibration Standards. A series of 12 standards covering a range

of 0-0.1 % FF A was prepared by gravimetric addition of palmitic acid to a refined and

deodorized soybean oil after it had been ron though an activated silica gel column to

59

Page 77: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

remove any traces of FF A and other oxygenated compounds. The FF A contents of the

standards were expressed in terms of % oleic acid.

Sample Preparation. Ten grams (± 0.001 g) of each oil sample or standard was

weighed on an analytical balance into a tared 15-mL clinical centrifuge tube. To the tube

containing the oil, 2 mL of the NaHNCN reagent solution was added using a calibrated

re-pipette. The tubes were capped, shaken on a vortex mixer for 30 s, and then

centrifuged for 5 min at 6000 rpm (-5000xg) to ensure consistent separation between the

oil and methanol phases.

Analytical Protocol. Approximately 1 ml of the methanol reagent was loaded into

the transmission flow cell and its single-beam spectrum recorded to serve as the

background spectrum. A new background spectrum was collected in the same manner

after every 20 samples or 1 h, whichever occurred tirst. For both samples and calibration

standards, 1 mL of the upper methanollayer formed after centrifugation was loaded into

the cell and its single-beam spectrum was recorded and ratioed against the

NaHNCNlMeOH background spectrum. The peak height of the carboxylate band at 1573

cm- l was then measured relative to an invariant baseline point at 1820 cm- l. The overall

sample preparation procedure and analytical protocol is illustrated in Figure 4.1.

Calibration and Validation. The calibration standards were taken through the

analytical protocol described above, and a calibration equation for the prediction of FF A

content was derived by plotting the concentrations of the standards (% oleic acid) vs. " , .' l 'I , ,

carboxylate peak height. The reproduCibÜity and accuracy of the FTIR method were

assessed by standard addition, spiking three different acid-free oils (canola, soybean and

sunflower) with known amounts ofpalmitic acid (w/w). These oils were each analyzed in

triplicate, on different days, by both the IUP AC titrimetric method (9) and the FTIR

method to allow a direct comparison of their performance. A comparative analysis of

locally purchased retined oils (soybean, sunflower, peanut, and corn oils and a

commercial oil blend) was also carried out.

4.4. RESULTS AND DISCUSSION

/""', The various approaches that have been investigated for the determination of FF A

content in edible oils by FTIR spectroscopy were reviewed in a previous paper (6), and

60

Page 78: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

-~---/

~ 2ml -MeOHlNaHNCN

~ 10g

1 ( Mix )

~

( Centrifuge ) FTIR 1 -(~) (Background) .........

l Spectrum 2 - Spectrum 1 (--A-) ./

FTIR 2 -(~) - (Sample)

Figure 4.1. Schematic diagram of the sample preparation and analytical protocol.

the limitations of each approach discussed. In that paper, these limitations were addressed

by the development of a new FTIR method based on previous work on AN determination

in lubricants, which employed the mild base potassium phthalimide as a signal­

transducing reagent i~ conjunction wiili differential spectroscopy to circumvent matrix

effects (6). Although the phthalimide FTIR procedure is bOth more accurate and

reproducible than conventional titration, it is ultimately not more sensitive per se. This is

largely due to the use of propanol as a solvent and polarity enhancer for the reaction,

effectively diluting the COO- IR signal. A means by which sensitivity could be improved

would be to treat the oil with methanol containing a base which is immiscible with the oïl

to facilitate the acid-base reaction as well as concentrate the FF A salts in the methanol

layer (14). Such a procedure would have the additional advantage of minimizing matrix

effects by partitioning out the spectral contribution of the oil. For highly accurate

analyses, a weak base would be required to avoid any saponification of the oil and one

61

Page 79: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

could use either the spectral changes associated with the loss of the base or formation of

FF A salts as a basis for quantitation. Examination of a range of reagents led to the

consideration ofthe sodium salt of carbodiimide (NaHNCN), which has a strong v (C=:N)

absorption at 2100 cm-l, is readily soluble in methanol, and is capable of converting

FF As to their carboxylate saIts but not capable of saponifying triacylglycerols. In the first

instance, the proportionate decrease in the intensity of the v (C=:N) band appeared to be a

very good measure of the amount of FF As spiked into oils. Although this reaction

worked, its was found that the reagent spectrum was unstable and that the band at 2100

cm-l slowly disappeared over a period of - four days with the concomitant appearance of

two new bands at 1650 and 1610 cm-l (Figure 4.2). Based on assignment of these two

bands to C=N stretching and NH bending vibrations, respectively, the structural

rearrangement from NaHN-C=:N -7 NaN=C=NH was postulated to be taking place in

solution over time. This conversion was conftrmed spectrally by dissolving NaHNCN in

methyl alcohol-d (MeOD), whereby the band at 1610 cm-l

shifted about 100 cm-l to

lower frequency owing to hydrogen-deuterium exchange.

0.30 B----"-1\

0.15 A

~ ~ = = ..c 0.00 ... 0 f'-l

..c <

-0.15

B

-0.30 2400 2200 2000 1800 1600

Wavenumber (cm-1)

Figure 4.2. Differentiai spectra for sodium hydrogen cyanamide in methanol. The spectra show the decomposition of the reagent over time. A = one day; B = four days.

62

Page 80: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

This molecular rearrangement, which was established to be complete within 4

days after preparation of the MeOH/NaHNCN reagent, did not affect its ability to convert

FF As to their respective salts without causing oil saponification. The reactivity of the

converted MeOHINaHNCN solution as weIl as its spectral characteristics remained

stable, with solutions kept at room temperature being used for up to two months without

any apparent deterioration in their efficacy. On the other hand, the measurement of the

. C=N band originaIly envisioned as a basis for quantitation is lost as a result of this

transformation. However, measurements made using the carboxylate band ofFFA salts at

1573 cm-I were both reproducible and responsive to low FFA levels because of the

concentration of the FF A salts in the methanollayer and the high extinction coefficient of

the carboxylate band. It was found that optimum sensitivity and reproducibility were

achieved with a 5:1 oil:methanol ratio (10 g oil plus 2 mL of "aged" MeOH/NaHNCN

reagent mixed in a standard 15-mL clinical centrifuge tube), producing consistent

reactions and reproducible separations of the MeOH and oil phases when centrifuged for

5 min at 5000 x g. Oil stability with respect to saponification by the reagent was assessed

by incubating oil samples with the reagent for 24 h and did not result in any measurable

oil hydrolysis, but did lead to a minor displacement effect (~0.01% FFA) due to sorne

additional oil migration into the methanollayer over extended periods oftime.

4.4.1. Calibration

A calibration curve was developed by using a set of standards covering the range

of 0.0 to 0.1 % FFA, prepared by adding palmitic acid to acid-free soybean oil. Figure 4.3

illustrates the differential spectra obtained for these standards using the optimized

analytical protocol outlined in Figure 4 1. The carboxylate anion produced and extracted

into the Me OH layer shows a rising absorbance at 1573 cm-l, the other spectral features

at 1650 cm- I and the split band covering 1760-1700 cm- I representing the v (C=N)

absorption of the base and the v (C=O) ester linkage absorption of a smaIl amount of

solubilized oil. A plot of the absorbance measured at 1573 cm -l, referenced to a single­

point baseline at 1820 cm-l, vs. % FF A (oleic acid) is presented in Figure 4.4. Linear

regression of the data obtained resulted in the following relationship:

63

Page 81: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

1573 cm-t

0.16

~ 0.12 CJ ::1

= .c .. 0 f'-.l

0.08 .c -<

1820 cm-1

0.04

1800 1750 1700 1650 1600 1550 1500

Wavenumber (cm-1)

Figure 4.3. DifferentiaI spectra for soybean oil spiked with palmitic acid (0.0 - 0.1%1 after carrying out the acid/base reaction. Spectra were recorded in a 100 J.lm cell at 8 cm­resolution.

, 'l,

, '

% FF A = 0.70292 A(1573/1820) - 0.00883

R2 = 0.9998; SD = 6.706xl0-4 [4.1]

The regression SD implies that FFA levels in the order of 1/1000th of a percent may be

measurable by this technique.

4.4.2. Validation

Validation and comparison of the FTIR method relative to the IUP AC titrimetric

method were carried out by analyzing three acid-free oils (soybean, canola and

sunflower) spiked with palmitic acid. Figure 4.5 presents comparative plots for triplicate

64

Page 82: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

0.16

,.... e 0.12 u

co N 00 ... ~ 1/') • e, ~ 0.08 CJ

== • ~ .c ,.. 0 fil .c 0.04 • -< t • t

•• 0.00

0.00 0.02 0.04 0.06 0.08 0.10

0/0 FFA w/w (Oleic Acid)

Figure 4.4. Calibration curve for %FF A in oïl obtained from the differential spectra in Figure 4.2. The %FF A is expressed as % oleic aeid (w/w). Error bar amplitude indicates mean ± SD of three replicates.

analyses of these oïls by FTIR spectroscopy and the IUP AC titration procedure,

respectively. The corresponding regression equations are:

, ' 1,/

R2 = 0.9998 [4.2]

IUPACFFA = 1.047 FFA + 2.3xlO-2

R2 = 0.9887 [4.3]

These results clearly indicate that the FTIR method tracks standard addition very

weIl, with a slope and an intercept very close to unity and zero, respectively, with an

overall error of about <0.001 %. The titrimetric plot clearly shows greater variability.

65

Page 83: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

0.12 -~ ~ 0.09 ~ C '-' ~ joooj

~ 0.06 ~ ~ .c < 0.03 ~ ~

0.00

0.12 -~ ~ ~ 0.09

= o .,.. ..... ~ !:: 0.06 .,.. ~ ~ .c < 0.03

~ 0.00

A

* I

• f • • .,;.

.. • •

0.00 0.02 0.04 0.06 0.08 0.10

FFA Added (%W/W)

B

0.00 0.02 0.04 0.06 0.08 0.10

FFA added (o/OW/w)

Figure 4.5. A plot of %FF A determined by the FTIR method (A) and %FF A determined by the titrimetric method (B) against the spiked amount. Error bar amplitude indicates mean ± SD of three replicates.

66

Page 84: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

.~ ..

Table 4.2 presents the data in terms of the relative mean difference (MD) and

standard deviation of the differences (SDD) for both accuracy (a) and reproducibility (r).

The MDa of 0.025% for the titrimetric method indicates a significant bias relative to

standard addition considering its SDDa, while the FTIR results are basically an order of

magnitude better in terms of both accuracy and reproducibility with no significant bias.

The apparent titrimetric over determination relative to the FTIR method may be due to

systematic errors associated with a slower acid/base reaction owing to the lower solvent

polarity, the substantive volume of solvent used, and C02 absorption, as well as

variability contributed by the visual endpoint determination. Table 4.3 presents

comparative triplicate FTIR and titrimetric data obtained for locally purchased samples of

refined oils. The two methods correlate (R2 = 0.83), showing a similar trend as the

standard addition data, with the titrimetric procedure predicting higher values

individually as well as an overall mean bias of ....Q.02% and yielding poorer

reproducibility.

Table 4.2. Mean Difference (MD) and Standard Deviation of the Differences (SDD) for Accuracy (a) for the Titrimetric and FTIR Procedures vs. Gravimetrie Addition and for Reproducibility (r) for Duplicate Analyses Carried Out by Titrimetric and FTIR Analyses, Respectively

Statistic FfIR Titration

MDa 5.84 X 10-4 2.50 X 10-2

" : 1',111 ,! f

SDDa 7.00 x 10-4 5.10 X 10-3

MDr 4.56 x 10-4 3.65 xlO-3

SDDr 9.47 x 10-4 9.86 X 10-3

The carbodiimide FTIR method was further examined in relation to expanding its

general utility by changing its oil:reagent ratio. By decreasing the oil:reagent ratio from

5:1 to 1:4, while maintaining the same carbodiimide concentration, a calibration covering

a range of 0-4% FFA was developed which produced the following calibration linear

regression equation:

67

Page 85: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

% FFA = 18.994*A(1573/1820) - 0.03598

R2 = 0.9999 SD = 0.0206 [4.4]

Thus simply by changing oil:reagent ratio, semi-refined and crude oils can also be

analyzed accurately using the same basic analytical protocol outlined in Figure 4.1. As

such, the method developed provides a simple FTIR-based analytical procedure for the

measurement of both very low and moderate levels of fatty acids in oils. Aside from

providing excellent sensitivity and reproducibility, it overcomes many of the limitations

associated with conventional titrimetric and potentiometric methods. There is no need to

use an FTIR spectrometer for this analysis per se, given that measurements at only two

wavelengths are required and a simple dual-wavelength fiUer instrument would suffice.

ln either case, the method is readily amenable to automation.

Table 4.3. Results of Triplicate Analyses of Locally Purchased Oil Samples by the IUP AC Reference (Titrimetric) Method and the FTIR Method

Titrimetric method FTIRmethod Oil type

%FFA(w/wt SD %FFA(w/w)a SD

Peanut oil 0.050 5.39xlO-3 0.030 7.03xl0-5

Sunflower oil 0.041 ., .," 6.40xl0-3 0.021 9.43xlO-4

Comoil 0.053 3.50xl0-3 0.024 4.39xlO-4

Mixed oil 0.061 7.15xl0-3 0.032 8.28xl0-4

Soybean oil 0.045 3.62xlO-3 0.020 1.62xlO-4

Overall Mean 0.050 0.026

a Expressed as % oleic acid

68

Page 86: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

I~'

4.5. CONCLUSION

Our work on FTIR analysis of lubricating oils led us to develop a novel approach

for the determination of acidity in non-aqueous systems based on the combined use of

signal transduction via a stoichiometric reaction and differential spectroscopy. In

subsequent work, we demonstrated the suitability of this approach for the quantitation of

FF As at levels of >0.2% oleic acid by employing the same reagent as used for the

determination of acidity in lubricating oils. In the present study, the use of sodium

carbodiimide and a modified procedure has allowed the analysis to be extended down to

FF A levels as low as 0.001 %, which would not be measurable by direct measurement of

FF A absorptions in the IR spectra of oils. With a simple adjustment in the reagent:oil

ratio the method becomes scalable and the analysis of semi-refined and crude oils

containing up to 4.0% FF A is possible. The strength of the method is its sensitivity and

allows for the possibility of using FF As as an indicator of lipid oxidation and to more

accurately monitor the refining processes carried out on edible oils.

4.6. ACKNOWLEDGMENT

F. R van de Voort acknowledges the financial support of the Natural Sciences

and Engineering Research Council of Canada and the cooperation of Canamera Foods for

providing oil samples for analysis. A AI-Alawi thanks Sultan Qaboos University for

financial support for his PhD studies.

4.7. REFERENCES , "

J" t

1. Haaland, D. M. Quantitative Infrared Analysis of Borosilicate Films usmg Multivariate Statistical Methods. Anal. Chem., 60, 1208-1217 (1988).

2. van de Voort, F. R, Ismail, A. A, Sedman, J., Dubois, J and Nicodemo, T. The Determination of Peroxide Value by Fourier Transform Infrared (FTIR) Spectroscopy. J. Am. Oil Chem. Soc., 71, 921-926 (1994).

3. Ma, K., Van De Voort, F. R, Sedman, J. and Ismail, A. A Stoichiometric Determination of Hydroperoxides in Fats and Oils by Fourier Transform Infrared Spectroscopy. J. Am. Oil Chem. Soc., 74, 897-906 (1997).

4. van de Voort, F. R., Sedman, J., Yaylayan, V. and Saint Laurent, C. Determination of Acid Number and Base Number in Lubricants by Fourier Transform Infrared Spectroscopy. Appl. Spectrosc., 57, 1425-1431 (2003).

69

Page 87: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

5.

6.

7.

8.

9.

10.

Il.

12.

13.

14.

15.

16.

17.

18.

19.

van de Voort, F. R, Sedman, J., Yaylayan, V., Saint Laurent, C. and Mucciardi, C. Quantitative Determination of Moisture in Lubricants by Fourier Transform Infrared Spectroscopy. Appl. Spectrosc., 58, 193-198 (2004).

AI-Alawi, A, van de Voort, F. R and Sedman, J. New Method for the Quantitative Determination of Free Fatty Acids in Oil by FTIR Spectroscopy. J Am. ail Chem. Soc., 81,441-446 (2004).

O'Brien, R D. Fats and ails: Formulating and Processing for Applications, Technomic Publishing Company, Lancaster, PA, 1998

Senorans, F. J. and Ibanez; E. Analysis of Fatty Acids in Foods by Supercritical Fluid Chromatography. Anal. Chim. Acta, 465,131-144 (2002).

International Union Pure and Applied Chemistry. Standard Methods for the Analysis if ails, Fats and Derivatives, th revised and enlarged edn., Paquot, c., and Hautfenne, A., Eds. Blackwell Scientific, Oxford, 1987

Official Methods and Recommended Practices of the American ail Chemists' Society, 4th edn., 1989,Champaign, IL. AOCS Official Method Ca-5a-40

Velasco, J., Andersen, M. L. and Skibsted, L. H. Evaluation of Oxidative Stability of Vegetable Oils by Monitoring the Tendency to Radical Formation. A Comparison of Electron Spin Resonance Spectroscopy with the Rancimat Method and DifferentiaI Scanning Calorimetry, Food Chem., 85, 623-632 (2004).

Nouros, P. G., Georgiou, C. A. and Polissiou, M. G. Automated Flow Injection Spectrophotometric Non-Aqueous Titrimetric Determination of the Free Fatty Acid Content of Olive Oil. Anal. Chim. Acta, 351(1-3),291-297 (1997).

Mariotti, E. and Mascini, M. Determination of Extra Virgin Olive Oil Acidity by FIA-Titration. Food Chem., 73, 235-238 (2001).

Kwon, D. Y. and Rhee, J. S. A Simple and Rapid Colorimetric Method for Determination of Free Fatty Acids for Lipase Assay. J Am. ail Chem. Soc., 63, 186-189 (198(». , " ,

Tur'yan, Va. 1., Bel'ezin, O. Yu; Kuselman, 1. and Shenhar, A. pH-Metric Determination of Acid Values in Vegetable Oils without Titration. J Am. ail Chem. Soc., 73, 295-301 (1996).

Takamura, K., Fuse, T., Arai, K. and Kusu, F. A Review of a New Voltammetric Method for Determining Acids. J Electroanal. Chem., 468, 53-63 (1999).

Dermaux, A., Sandra, P. and Ferraz, V. Analysis of Free Fatty Acids and Fatty Acid Phenacyl Esters in Vegetable Oils and Margarine by Capillary Electro­Chromatography. Electrophoresis, 20, 74-79 (1999).

Kotani, A., Kusu, F. and Takamura, K. New Electrochemical Detection Method in High-Performance Liquid Chromatography for Determining Free Fatty Acids. Anal. Chim. Acta, 465, 199-206 (2002).

Rosenfeld, J. M. Application of Analytical Derivatizations to the Quantitative and Qualitative Determination of Fatty Acids. Anal. Chim. Acta, 465, 93-100 (2002).

70

Page 88: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

20. Senorans, F. J. and Ibanez, E. Analysis of Fatty Acids in Foods by Supercritical Fluid Chromatography. Anal. Chim. Acta, 465, 131-144 (2002).

21. Lanser, A. C., List, G. R, Holloway, R K. and Mounts, T. L. FTIR Estimation of Free Fatty Acid Content in Crude Oils Extracted from Damaged Soybeans, J. Am. Oil Chem. Soc., 68, 448-449 (1991).

22. Ismail, A. A., van de Voort, F. R and Sedman, J. Rapid Quantitative Determination of Free Fatty Acids in Fats and Oils by FTIR Spectroscopy. J. Am. Oil Chem. Soc., 70, 335-341 (1993).

23. Bertran, E., Blanco, M., Coello, J., Iturriaga, H., Maspoch, S. and Montoliu, 1. Determination of Olive Oil Free Fatty Acid by Fourier Transform Infrared Spectroscopy, J. Am. Oil Chem. Soc., 76, 611-616 (1999).

24. Man, Y. B., Moh, M. H. and van de Voort, F. R Determination of Free Fatty Acids in Crude PaIm Oil and Refined-Bleached-Deodorized Palm Olein Using Fourier Transform Infrared Spectroscopy, J. Am. Oil Chem. Soc., 76, 485-490 (1999).

25. Verleyen, T., Verhe, R, Cano, A., Huyghebaert, A. and De Greyt, W. Influence of Triacylglycerol Characteristics on the Determination of Free Fatty Acids in Vegetable Oils by Fourier Transform Infrared Spectroscopy, J. Am. Oil Chem. Soc., 78, 981-984 (2001).

26. Zhang, H-Z and Lee, T-C. Rapid Near-Infrared Spectroscopic Method for the Determination of Free Fatty Acid in Fish and Its Application in Fish Quality Assessment. J. Agric. Food Chem., 45, 3515-3521 (1997).

27. Man, Y. B. C. and Moh, M. H. Determination of Free Fatty Acids in Paim Oil by Near-Infrared Reflectance Spectroscopy, J. Am. Oil Chem. Soc., 75, 557-562 (1998).

28. Muik, B., Lendl, B., Molina-Diaz, A. and Ayora-Canada, M. J. Direct, Reagent­Free Determination ofFree Fatty Acid Content in Olive Oil and Olives by Fourier Transform Raman Sp~ctrop1etry~ :Anal. Chim. Acta, 487, 211-220 (2003).

29. Canada, M., Medina, A. and Lendl, B. Determination of Free Fatty Acids in Edible Oils by Continuous-Flow Analysis with FT-IR Spectroscopic Detection, Appl. Spectrosc., 55, 356-360 (2001).

71

Page 89: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

BRIDGE

In Chapter 4, an approach to FF A analysis was developed based on the treatment

of oils with an oil immiscible solvent containing a weak base to convert the FF A to their

carboxylate salts and concentrate the salts in a small volume to increase the sensitivity of

the method. By comparison with the method described in Chapter 3, this approach also

provided the benefit of eliminating the need to prepare two samples to obtain one

analytical result as well as allowing an overall simplification of the analysis. This

extraction/concentration approach is considered in Chapter 5 as a possible means by

which the moisture content of edible oils can be determined.

72

Page 90: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

CHAPTER5

A NEW FTIR METHOD FOR THE DETERMINATION OF MOISTURE IN

EDIBLE OILS

5.1. ABSTRACT

A rapid, practical, and accurate FTIR method for the determination of moisture

content in edible oils has been developed based on the extraction of water from oil

samples into dry acetonitrile. A calibration curve covering a moi sture content range of 0-

2000 ppm was developed by recording the mid-IR spectra of moi sture standards,

prepared by gravimetric addition of water to acetonitrile that had been dried over

molecular sieves, in a 500 J.1m ZnSe transmission flow cell and ratioing these spectra

against that of the dry acetonitrile. Water was measured in the resulting differential

spectra using either the OH stretching (3629 cm-1) or bending (1631 cm-1

) bands to

produce linear standard curves having SDs of - ±20 ppm. For moi sture analysis in oils,

the oil sample was mixed with dry acetonitrile in al: 1 w/v ratio, and after centrifugation

to separate the phases, the spectrum of the upper acetonitrile layer was collected and

ratioed against the spectrum of the dry acetonitrile used for extraction. The method was

validated by standard addition experiments with samples of various oil types, as well as

with oil samples deliberately contaminated with alcohols, hydroperoxides, and free fatty

acids to investigate possible interferences from minor constituents that may be present in

oils and are potentially extractable ,illtO !.acetonitrile. The results of these experiments " ,

confirmed that the moisture content of edible oils can be assessed with high accuracy (on

the order of ±1O ppm) by this method, thus providing an alternative to the conventional,

but problematic Karl Fischer method and facilitating the routine analysis of edible oils for

moi sture content.

5.2. INTRODUCTION

Moisture content is an important parameter associated with the processing and

quality of edible oils. Commonly used in pro cesses such as degumming and refining,

water is subsequently removed by centrifugation, adsorption or vacuum drying. Although

water has limited solubility in fats and oils, the moi sture content must be minimized for

73

Page 91: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

the effective adsorption of remaining soap traces after retining and to reduce lipid

hydrolysis during and after processing (1) GeneraIly, the moi sture content ofretined oils

should be <0.1% (1000 ppm) and preferably <500 ppm (2). Methods for moi sture

ap.alysis in edible oils that serve as Official Methods of the American Oil Chemists'

Society (AOCS) (3) include evaporative, distillation or titrimetric procedures as weIl as

the Karl Fischer (KF) method, which is based on the selective reaction of water with a

mixture of iodine, a base, sulfur dioxide, and an alcohol followed by either titrimetric,

potentiometric (4,5) or coulometric (6,7) quantitation. The coulometric KF method is

considered to be the most accurate and sensitive method available and is the gold

standard for the determination of a wide range of moi sture contents (1-25,000 ppm).

However, it is an exacting procedure, uses expensive and environmentally problematic

reagents, and is affected by oil oxidation end products such as aldehydes, ketones, and

hydroperoxides, as these can undergo aldol condensation and/or redox reactions under the

conditions of the test (8). Thus, there is effectively no simple, convenient, sensitive

method available for the analysis of moisture in fats and oils.

At tirst glance, IR spectroscopy would appear to be an obvious instrumental

approach for the direct measurement of moisture in edible oils, given the strong

absorption bands of water in the mid-IR portion of the spectrum, (9) and a limited

feasibility study of the determination of moisture content in crude palm oil by FTIR

spectroscopy has recently been reported (10) However, a number of complicating factors

that were not taken 'into account in: the' previous study must be considered in the

development of a generalized FTIR method for the determination of moi sture in edible

oils. First of aIl, the band shapes and intensities of the water absorptions, particularly the

OH stretching vibrations, are strongly affected by the extent of hydrogen bonding

between water molecules, which will depend on the moisture content in an oil, as weIl as

between water and other hydrogen-bonding constituents. Furthermore, OH-containing

constituents that are frequently present in oils (e.g., alcohols, free fatty acids (FFAs),

and/or hydroperoxides) not only perturb the water absorptions via hydrogen-bonding

interactions but also give rise to spectrally interfering bands. In a recent publication (11),

we reported a new approach to the FTIR analysis of moisture in new and used mineral­

based lubricants in whlch similar "matrix effects" were overcome by the combined use of

74

Page 92: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

/-the stoichiometric reaction of water with dimethoxypropane (DMP) to produce acetone

(12,13) and differential spectroscopic techniques to quantitate this end product by

measurement of its v (C=O) absorption. This approach, however, is not applicable to

edible oils because the triacylglycerol ester linkages give rise to an intense v (C=O)

absorption that would completely mask the v (C=O) band of acetone in a transmission

spectrum, except at extremely short path lengths. Moreover, the use of ATR to overcome

this limitation is impractical owing to the difficulty of preventing the sample from

picking up moisture from the environment, nor would it provide sufficient sensitivity for

the analysis of moi sture at the low levels present in edible oils. Hence, an alternative

approach is required for edible oils, and this paper describes the evaluation of an

approach based on the extraction of water from the oil into a suitable solvent, culminating

in the development of an acetonitrile extraction method that allows for the accurate

determination of low levels of moisture in edible oils.

5.3. METHODOLOGY

Reagents. HPLC grade acetonitrile, obtained from Fisher Scientific (S~. Louis,

MO), was kept over 4-8 mesh 4A molecular sieves and dispensed using a re-pipette

(Hirschmann-Laborgerate, Germany) protected by desiccant to prevent moisture ingress.

Various refined edible oils were obtained locally, and sorne samples of these oils were

spiked with constituents spectrally representative of contaminants found in edible oils;

these included tert-butyl hydroperoxide, glycerol, and oleic acid, aIl obtained from

Sigma-Aldrich.

Instrumentation. The FTIR spectrometer used for this study was a Bornem

WorkIR (Bornem, Quebec City, PQ, Canada) equipped with a DTGS detector and purged

with dry air using a Balston dryer (Balston, Lexington, MA). The spectrometer was

controlled by an IBM-compatible Pentium 150-MHz PC running proprietary Windows­

based UMPIRE® (Universal Method Platform for InfraRed Evaluation) software

(Thermal-Lube, Pointe-Claire, PQ, Canada). A 500 J.lm ZnSe transmission flow cell

(International Crystal Laboratories, Garfield, NJ) equipped with Luer fittings was

employed for the handling of oil samples. As illustrated in Figure 5.1, the cell inlet was

connected to a 10 cm, 18-gauge stainless steel aspiration needle via flexible silicone

75

Page 93: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

tubing, and the outlet line was connected to vacuum via a trap and was titted with a valve

to allow both aspiration of the sample through the cell and emptying of the celI; in the

latter process, the air entering the cell was passed through a desiccant tube containing

molecular sieves and Drierite to prevent environmental moisture from contaminating the

analytical system. AlI spectra were collected by co-adding 32 scans at a resolution of 8

cm-1 and a gain of 1.0.

Desiccant

Needle

CellHolder

Valve ~~~o-__ ~Vacuum

--

Acetonitrile layer

Oillayer

Cell Holder Base

Figure 5.1. Schematic diagram of the FTIR sample handling system. To facilitate loading, the upper acetonitrile layer of the centrifuged or separated sample (1) is vacuum aspirated into the IR cell using the 10 cm stainless steel needle attached by flexible tubing to the IR cell. Between sample loadings, the needle is inserted into the desiccant tube outlet (2) to ensurethat air used to flush the tubing and the cell does not introduce any moisture.

Analytical Protocol. The sample preparation procedure and analytical protocol

developed for the determination of the moisture content of oils by extraction of the

moisture into dried acetonitrile and its quantitation by FTIR spectroscopy are illustrated

76

Page 94: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

in Figure 5.2. For the analysis of oïl sampi es having moisture contents in the range of 50-

2000 ppm, a 1:1 (v/w) acetonitrile:oil ratio was established to be suitable. Accordingly, 5

g of the sample was added to a tared 15-mL clinical centrifuge tube and weighed on an

analytical balance, with the weight recorded to ±0.001 g, and 5 mL of the dried

acetonitrïle was then added to the tube using a calibrated re-pipette. The tubes were

capped, shaken on a vortex mixer for 30 s, and then centrifuged for 2 min at -6000 rpm

(-5000xg) to separate the oïl and acetonitrile phases. Altematively, adequate separation

of the phases could be achieved by letting the sample stand for -10 min, eliminating the

need for the centrifugation step. Approximately 2 mL of the upper acetonitrile layer was

then aspirated into the transmission flow ceIl, and its spectrum recorded and ratioed

against the spectrum of the dry acetonitrile used to extract moisture from the oïl samples.

A new acetonitrile background spectrum was collected after every 20 samples or after 1

h, whichever came first.

~~ ~ 5 g Acetonitrile

1iI~ -! !

1 Mix 1 ~--+IAJLI !ëentrifugel \t .--------,

l Spectrum 2 - Spectrum 1 = 1 ~ 1

? -+ f~3 --+I~I

Figure 5.2. Schematic diagram of the sample preparation procedure for moi sture analysis of edible oils and the FTIR spectral analytical protocol.

77

Page 95: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

Calibration and Validation. Two calibration approaches were evaluated: one based

on the use of primary water/acetonitrile standards and the other on water/oil standards.

Twenty primary water/acetonitrile standards were prepared gravimetrically by adding

distilled water to dry acetonitrile to coyer a range of 0-1000 ppm, and a set of 12

water/oil standards (0-2000 ppm) were similarly prepared by gravimetrically adding

distilled water to canola oïl previously kept over molecular sieves for a minimum of 1

week. The primary standards were analyzed directly by FTIR spectroscopy, while the

water/oïl standards were taken through the analytical protocol illustrated in Figure 5.2. In

both cases, calibration equations were developed relating the moi sture added (in ppm) to

the three measurable water bands at 3629, 3541, and 1631 cm-l, relative to a baseline

point at 2500 cm-l. The reproducibility and accuracy of the FTIR method were assessed

by standard addition of water to various oils (olive, corn, safflower, peanut, and

sunflower), each of these samples being analyzed in triplicate, on different days, with the

FTIR-predicted moi sture contents being compared with the amounts added.

5.4. RESULTS AND DISCUSSION

5.4.1. General Considerations

The objective of this study was to develop a generally applicable method for the

routine analysis of moisture in edible oils based on extraction of the moisture into a

suitable solvent followed by quantitation by FTIR spectroscopy. The solvent initially

considered was 020, in wbich each mqieculeofH20 extracted from the oil sample would "

rapidly undergo hydrogen-deuterium exchange to produce two HOO molecules, thereby

effectively doubling the spectral signal for each molecule of H20 present. Thus, in

princip le, samples could be analyzed by mixing a defined excess of 0 20 with the oil,

allowing the oïl and 020 to separate, and quantitating the HOO formed by measuring the

intensity of the characteristic HOO bands at 3375 and 1450 cm-l in the spectrum of the

0 20 layer. However, although tbis approach worked well for clean refined oils, the

presence of OH-containing constituents in the oil (e.g., alcohols, hydroperoxides or

FFAs), which also undergo hydrogen-deuterium exchange with 0 20, albeit more slowly,

resulted in significant overestimation of moisture content, as evidenced by standard

addition experiments. This problem, in conjunction with the relatively high cost of D20,

78

Page 96: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

led us to seek another extraction solvent in which the concentration of H20 could be

accurately quantitated.

Among the solvents investigated, acetonitrile was ultimately selected. It was

found to be sufficiently polar to be immiscible with edible oils and solubilize water while

having limited capacity to solubilize the other potentially interfering OH-containing

constituents that may be present in oils. Acetonitrile is also a highly suitable solvent in

which to measure moi sture levels by IR spectroscopy as it does not absorb strongly in the

portions of the mid-IR spectrum where water absorbs. This lack of spectrally interfering

bands makes it possible to use path lengths of up to 1000 J.lm for the analysis of low

levels of moisture, thereby providing high sensitivity. Finally, in contrast to the spectrum

of water itself, in which the-two OH stretching absorptions are blended into one large

diffuse band owing to the hydrogen-bonding network, the spectrum of water diluted in

acetonitrile at mole fractions of <0.1 exhibits two discrete OH bands, which has been

attributed to the binding of water molecules exclusively to acetonitrile rather than to each

other at these high dilutions (14). Thus, the spectra ofwater added to acetonitrile at levels

of 400 and 800 ppm, presented in Figure 5.3, exhibit clearly delineated bands at 3629

and 3541 cm"l together with the weaker HOH bending vibration at 1631 cm"l.

1.0

1631 cm"I_--a

0.8 CIJ CJ

= 0.6 = ,.Q &. Q rI.l 0.4 ,.Q

-< 0.2

0.0 r i

3600 2000 1600

Wavenumber (cm-1)

Figure 5.3. Spectra ofwater/acetonitrile standards (0, 400, and 800 ppm added water) (upper series) and the corresponding speCtra obtained after subtraction of the spectrum of the acetonitrile employed to prepare the standards (lower series).

79

Page 97: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

~'. (

5.4.2. Calibration Development

Primary calibrations were devised using 20 standards prepared by addition of

water to acetonitrile (dried over molecular sieves) to obtain moi sture contents in the range

of 0-1000 ppm based on the gravimetrically added amount. The contribution of any water

present in the acetonitrile was not taken into account in calculating the moi sture contents

of these standards because the method was designed to employ differential spectra,

obtained by subtraction of the spectrum of the acetonitrile used to prepare the standards

from the spectrum of each standard, thereby eliminating the spectral contributions of any

residual water in the dried acetonitrile.

The spectra of the standards were recorded in a 500 Jlm celI, and the following

linear regression equations wère obtained for the 3629, 3541, and 1631 cm-1 bands in the

differential spectra:

H20 (ppm) = 3420.82* A 3629 cm-1 - 4.29

R2 =0.999 SD = 17.7 [5.1]

H20 (ppm) = 3934.92* A 3541 cm-! - 5.55

R2 =0.999 SD = 17.4 [5.2]

H20 (ppm) = 6941.50*A 1631 cm-1 + 2.33

SD = 16.9 [5.3]

These equations indicate relative sensitivities of -3.4, 3.9, and 6.9 ppm per

milliabsorbance unit for the 3629, 3541, and 1631 cm-! bands, respectively, with SDs of

<20ppm.

Figure 5.4 presents a series of differential spectra obtained for a second set of

standards (n = 12) prepared by spiking canola oH with water (0-2000 ppm) and

processing the samples in accordance with the protocol illustrated in Figure 5 2. The

calibration equations derived from these spectra were also linear, with similar slopes,

correlation coefficients, and regression SDs as those obtained using the primary

80

Page 98: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

r-..

water/acetonitrile standards, but had negative intercepts of ~280 ppm, indicating that the

canola oil contained about 0.03% moisture. A notable difference between the differential

spectra of the two sets of standards (Figure 5.3 vs. Figure 5.4) is the presence of a

carbonyl band (peak D) at 1740 cm- l in the spectra of the set prepared by extraction of

water from oil into acetonitrile. It was detennined experimentally that the positions of the

v (C=O) absorptions of oils (triacylglycerol ester linkage) and FFA mixtures derived from

the same oils were identical in acetonitrile. However, although FFAs are freely soluble in

acetonitrile, it was considered unlikely that peak D was solely due to FF As extracted

from the oïl, owing to their low levels in refined oils «0.05%), and this was borne out by

further examination of the spectra, in that peak D was not accompanied by any FF A

absorption at ...... 3300 cm- l. Thus, the observation ofpeak D was taken as an indication that

sorne amount of oil was extracted into the acetonitrile layer.

Figure 5.4. Differentiai spectra obtained after extraction of water from water/canola oil standards (0-2000 ppm water) into acetonitrile, recording the spectrum of the acetonitrile layer, and ratioing the spectrum against that of the acetonitrile extraction solvent. The major spectral features are identified as follows: A and B (3629 and 3541 cm-l, respectively), OH stretching vibrations; C, water association band; D (1740 cm- l

),

ester carbonyl band of oil extracted into acetonitrile; E (1631 cm-\ HOH bending vibration.

81

Page 99: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

Accordingly, the miscibility of various oils with acetonitrile was assessed by

mixing the oil with acetonitrile in al: 1 (w/v) ratio and measuring the height of the

carbonyl band in the acetonitrile spectrum after phase separation was complete. The oils

were detennined to be slightly miscible (-0.5%) to an extent that varied somewhat with

the overall degree of unsaturation, which could mean that moi sture detennination using

the primary (water/acetonitrile) calibration could be affected by a dilution error that

would depend on the type of oil analyzed. To investigate this possibility, five oils (olive,

corn, safflower, peanut, and sunflower), all pre-dried over molecular sieves, were each

spiked with three levels ofmoisture (approximately 100, 300, and 500 ppm) and analyzed

in duplicate in accordance with the analytical protocol depicted in Figure 5.4. Regression

(forced through the origin) of the means of the duplicates against the amount of moisture

gravimetrically added to all the oils produced the following Z-reg relationship:

Predicted H20 = 0.986 * Added H20

R2 = 0.998 SD = 8.1 ppm [5.4]

Based on this experimental evidence, the use of calibrations based on simple

water/acetonitrile standards was considered a suitable means by which to quantify

moisture in edible oils.

5.4.3. Consideration of Interfering Con$tituents

Many minor constituents that may be present· in oils can be extracted into

acetonitrile and hence are potential sources of spectral interference in the quantitation of

moisture by the method developed in the present study. Of particular concern are OH­

containing molecules such as FF As, hydroperoxides, and alcohols (glycerol, mono- and

diglycerides). Indeed, as illustrated in Figure 5.5, which shows the overlaid differential

spectra of acetonitrile spiked with glycerol, hexanol, tert-butyl hydroperoxide, oleic acid,

and water, most of these constituents have bands that overlap significantly with one of the

two OH stretching bands (3541 cm-1) employed for quantitation of moisture. However,

the higher-frequency OH stretching band (3629 cm-1) is largely in the clear.

82

Page 100: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

Although the HOH bending band (at 1631 cm-1) does not suffer from

interferences from OH-containing constituents, the use of this band for quantitation of

moi sture is not optimal because of the resulting ~2-fold reduction in sensitivity (cf. Eqs.

[1]-[3]).

A 0.6

0.4

0.2

0.0 ....... --

3600 3400 3200 3000

Wavenumber {cm-Il

,t,

" . Figure 5.5. Differentiai spectra of acetonitrile spiked with hexanol (A), glycerol (B), tert-butyl hydroperoxide (C), oleic acid (0), and water (shaded). Note: The spectra are not on the same scale.

The effects that OH-containing constituents present in an oil would have on the

accuracy of the method was evaluated quantitatively by analyzing 24 standards prepared

gravimetrically by spiking corn oil with water (0-1000 ppm) and further spiking portions

of each of these standards with (i) oleic acid (0.09-0.8% w/w), (ii) glycerol (0.05-0.1 %

w/w) , or (iii) tert-butyl hydroperoxide (0.1-1 % w/w). Comparison of the moisture

predictions for the 18 standards in subsets (i), (ii), and (iii) with those for the 6 standards

spiked only with water demonstrated that the predicted values obtained using the 3541

83

Page 101: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

cm-1 band were high, roughly in proportion to the amount of "contaminant" added; the

largest effect was observed for glycerol, as would be expected given the extensive band

overlap illustrated in Figure 5.5, followed by the hydroperoxide and then the FF A. On

the other hand, the predictions of moi sture content based on either the 3629 or the 1631

cm-1 band showed no such effects. Regression of the predicted moi sture contents for an

additional subset, prepared by spiking four water/oil standards with randomized amounts

of all three contaminants, against the "expected" values (calculated from the predictions

for the standards by correction for the dilution due to spiking) yielded the following Z-reg

equations for the 3629 and 1631 cm-1 bands, respectively:

Predicted H20 3629 cm-1

= 1.054 * "Expected" H20

R2 =0.998 SD = 11.9ppm [5.5]

Predicted H20 1631 cm-1 = 1.050 * "Expected" H20

R2 =0.997 SD = 18.3 ppm [5.6]

These results éonfirmed that peak height measurements at 3629 or 1631 cm-1

allowed for the accurate quantitation ofmoisture in oils in the presence ofOH-containing

constituents at the levels commonly associated with edible oils.

5.4.4. Factors Affecting A~alytical Accu'tacy

Owing to the ubiquitous nature of moisture in the environment, precautions were

required to avoid moisture contamination of the sample handling system A simple, but

effective means of minimizing moi sture contamination was devised, as described in the

Experimental section and depicted in Figure 51, without which reproducible calibrations

over the range of 0-200 ppm could not be obtained. Above 200 ppm, the effects of

environmental moisture were relatively minor; however, to cover aIl contingencies, this

sample handling system was used routinely. Furthermore, in experiments conducted to

determine the relative degree of analytical stability ànd reproducibility when samples

were analyzed over a period of three days, deterioration of the analytical performance as

. time progressed was observed and eventually attributed to slow adsorption of water on

84

Page 102: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

the walls of glass centrifuge tubes. This was confirmed by the finding that reproducible

results were obtained in similar experiments when plastic centrifuge tubes fitted with

septa were employed in place of capped glass tubes. The results of these experiments not

only convey sorne of the precautions that must be taken to prevent loss or pickup of )

moi sture when analyzing low levels of moisture in oil samples but also serve to illustrate

the sensitivity of the FTIR method.

With proper precautions, very accurate data (on the order of ±10 ppm) can be

obtained, based on standard addition experiments. For maximum sensitivity, the 3629 cmc

1 OH stretch is the preferred measurement; however, the 1631 cm-l band is useful for

higher moi sture levels (above 2000 ppm). Although not detailed in this paper, additional

sensitivity (down to ~2 ppm) can be attained by increasing the pathlength (upper limit

~ 1 000 ~m) or using a 2: 1 oil:acetonitrile ratio in the extraction step, or both. However, at

such low moisture levels, the main limitation is not the sensitivity of the spectroscopic

measurement per se, but rather the difficulty of ensuring against moi sture pickup or loss,

especially after extraction and during analysis.

5.5. ACKNOWLEDGEMENTS

The authors would like to thank Sultan Qaboos University, Oman, and the Natural

Sciences and Engineering Research Council (NSERC) of Canada, for their financial

support of this research. Thanks to Charles Chaney for suggestions related to this work

and to Sophie Archelas for her technical assistance.

5.6. REFERENCES

1. O'Brien, R. D., Fats and Oils: Formulating and Processing for Applications, Technomic Publishing Company, Lancaster, PA, 1998.

2. Rossell, J. B. "Classical Analysis of Oils and Fats", in Analysis of Oils and Fats, R. J. Hamilton and J. B. Rossell, Eds., Elsevier Applied Science Publisher Ltd., London, England, 1986, Chap. 1, p. 52.

3. Official Methods and Recommended Practices of the American Oil Chemists' Society (AOCS, Champaign, IL, 1989), 4th edn.

4. Cedergren, A. and Luan, L. Potentiometric Determination of Water Using Spent Imidazole-Buffered Karl Fischer Reagents. Anal. Chem., 70, 2174-2180 (1998).

85

Page 103: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

5. Rosvall, M., Lundmark, Land Cedergren, A. Computer-Controlled, Coulometric Karl Fischer System for Continuous Titration of Water Based on Zero CUITent Potentiometry. Anal. Chem., 70, 5332-5338 (1998).

6. Cedergren, A and Jonsson, S. Diaphragm-Free Cell for Trace Determination of Water Based on the Karl Fischer Reaction using Continuous Coulometric Titration. Anal. Chem., 69, 3100-3108 (1997).

7. Cedergren, A. and Jonsson, S. Progress in Karl Fischer Coulometry using Diaphragm-Free Cells. Anal. Chem., 73, 5611-5615 (2001).

8. Cedergren, A. Reaction Rates between Water and the Karl Fischer Reagent. Ta/anta, 25, 256-271 (1978).

9. Smith, B. Infrared Spectral Interpretation: A Systematic Approach (CRC Press, Boca Raton, FL, 1999), Chap. 1, p. 22.

10. Man, Y. B. C and Mirghani, M. E. S. Rapid Method for Determining Moisture content in Crude Palm Oil by Fourier Transform Infrared Spectroscopy. J. Am. Oil Chem. Soc., 77, 631-637 (2000).

11. van de Voort, F. R., Sedman, J., Yaylayan, V., Saint Laurent, C. and Mucciardi, C. Quantitative Determination of Moisture in Lubricants by Fourier Transform Infrared Spectroscopy. Appl. Spectrosc., 58, 193-198 (2004).

12. Critchfield, F. E. and Bishop, E. T. Water Determination by Reaction with 2, 2-Dimethoxypropane. Anal. Chem., 33, 1034-1035 (1961).

13. Dix, K. D., Sakkinen, P. A., and Fritz, J. S. Gas Chromatographie Determination of Water Using 2, 2-Dimethoxypropane and a Solid Acid Catalyst. Anal. Chem., 61, 1325-1327 (1989).

14. Le Narvor, A., Gentric, E. and Saumagne, P. Study of Binary Mixtures of Water with Severa! Proton Acceptors by Infrared Spectroscopy in the Region of OH Stretching Bands. Cano J. Chem., 49, 1933-1939 (1971).

86

Page 104: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

BRIDGE

The methods developed in Chapters 4 and 5 share a common analytical approach,

namely the extraction of a constituent of interest from an edible oU into an oU immiscible

solvent and its spectroscopic quantification. The development of these sensitive and

accurate methods is only part of the solution in an industry that is looking to maximize

efficiency and reduce costs through automation. Because the extraction based procedures

use low viscosity solvents, they are amenable to being implemented on an automated

system equipped with an auto-sampIero Chapter 6 describes the implementation of

algorithms and procedures as well as validation of the methodology developed to effect

the automated analysis of FF A and moi sture in edible oils using such a system based on

the methods described in Chapters 4 and 5.

" ,

87

Page 105: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

CHAPTER6

AUTOMATED FTIR ANALYSIS OF FREE FATTY ACIDS OR MOISTURE IN

EDIBLEOILS

6.1. ABSTRACT

An FTIR spectrometer coupled to an auto-sampler and attendant methodologies

for high-volume automated quantitative analysis of free fatty acids (FF A) and moi sture in

edible oils are described. Samples are prepared by adding 20 g of oil to a 50 ml screw­

capped vial, to which Îs added either a methanol/NaHNCN solution or dry acetonitrile in

a 1: 1 (w/v) ratio for FF A or H20 analysis, respectively. After capping with Mylar-lined

septum caps, the vials are loaded into an auto-sampler tray, which is then agitated

vigorously to extract the constituent of interest from the oil into the oil-immiscible

solvent, and are then left to stand for ~ 10 min to allow for phase separation. The upper

solvent layer in each vial is aspirated successively into the IR ceIl, with the Mylar seal

allowing facile auto-sampler needle penetration into the vials. The spectra of the sample

extraction solvents serve as spectral backgrounds as weIl being used to monitor cell path

length and verify cell loading. FF A and H20 analyses are carried out using 100 /-lm and

500 /-lm CaF2 ceIls, respectively. For FFA analysis, quantification is achieved using the

v (COO-) band at 1573 cm-l, while moisture is determined using water absorption bands

at 3541 cm-l or 1631 cm-l, depending on the moisture range of the samples. Calibration

procedures and data are presented. The spectrometer and auto-sampler are controlled

using proprietary UMPlRE Pro® (Universal Method Platform for Infra Red Evaluation)

software, which provides a simple user interface and automates the spectral analysis; the

output data can also be sent to a Laboratory Information Management System (UMS).

Validation and performance data obtained with this automated system demonstrate that it

is capable of analyzing >60 samples/h, a rate commensurate with the throughput required

by commercial contract or high-volume process controllaboratories.

6~2. INTRODUCTION

Edible oils represent one of the primary constituents required for the formulation

and manufacture ofproducts by the food industry. Oils are extracted from a wide range of

88

Page 106: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

raw materials and generally undergo relatively standardized extraction and refining

processes and supplemental transformations such as fractionation, hydrogenation and/or

interesterification prior to use. During processing, a variety of physico-chemical oïl

quality parameters are monitored, including the determination of free fatty acids (FF A),

moisture (H20), peroxide value, iodine value, and saponification number, to name just a

few1• Typically, QC laboratories have been limited to using standardized (AOCS2 or

IUP AC3) manual wet chemical methods to monitor oïls in process, requiring a variety of

analytical setups and multiple reagents as well as skilled personnel to carry them out,

although these methods have been automated to sorne extent through the use of automatic

titrators and flow injection systems.

The Mc Gill IR Group has worked toward developing Fourier transform infrared

(FTIR) spectroscopy as a rapid instrumental technique for the routine analysis of edible

oils. Methodologies have been deve10ped for the determination of FF A 4,5, saponification

number6, iodine value6

,7, trans and cis content',9, ansidine valuelO, solid fat index 11 , and

peroxide valueI2-14

, amongst others. From the standpoint of implementation, most of

these methods are semi-automated, with the software providing a simple user interface as

well controlling the instrument, performing all calculations, and storing the results

automatically. These FTIR methods provide improved sample turnaround and a reduction

in reagent use relative to wet chemical methods. However, they do not provide the leve1

of speed required by contract or centralized QC laboratories handling more than 100

samples/d, the key factor limiting sample throughput being the high viscosity of edible

oils. On the other hand, in the case of new FTIR methods for the determin'ation of FF A 15

and H2016 in edible oils that have recently been deve10ped in our laboratory, this

limitation is obviated because the constituent of interest is extracted into a low-viscosity

solvent prior to FTIR analysis. Indeed, these new FTIR methods have the potential for

much higher sample throughputs than auto-titration systems, which presently provide the

most rapid means of obtaining reliable FF A or moisture data. In addition to reducing the

sample-flow constraints associated with high-viscosity oils, these extraction-based

methods largely e1iminate problematic matrix effects and provide much greater

sensitivity than can generally be achieved by direct FTIR analysis of oils. This paper

describes the instrumentation, analytical protocols, and software associated with an

89

Page 107: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

automated FTIRIauto-sampler system designed to analyze for either FF A or H20 in

edible oils at rates ofup to 100 samples per hour.

6.3. METHODOLOGY

Instrumentation. The analytical system used couples a Bornem WorkIR (Bornem,

PQ, Canada) FTIR spectrometer with a Model 223 Gilson auto-sampler (Gilson, Inc.,

Middleton, WI) and a positive displacement micro-pump (Vissers Sales Corp., ,Aurora,

Ont., Canada). The spectrometer is equipped with a DTGS detector and uses 100 and 500

~m CaF2 transmission flow cells (International Crystal Laboratories, Garfield, NJ, USA)

for FF A and H20 analysis, respectively, with Luer-Lock cell connectors used to facilitate

cell changes.

Software. The spectrometer, micro-pump and auto-sampler were all controlled by

an IBM-compatible PC running proprietary Windows-based UMPlRE® Pro (Universal

Method Platform for InfraRed Evaluation) software, developed by Thermal-Lube Inc.

(Pointe-Claire, PQ, Canada). Built on the Microsoft.NET Framework, UMPlRED Pro

software is designed to work with Windows XP or Windows 2000 operating systems.

UMPlRE Pro is a self-contained platform that allows for the acquisition of spectra,

analysis of spectral data, and cataloguing of results without operator intervention. For

systems equipped with an autosampler and pump (such as that used in the present work),

the analysis of all the samples loaded into the autosampler tray is fully automated. All

parameters àssociated with the three fundamental tasks controlled by the software

(spectral acquisition, data processing, and cataloguing of results) are predefined for each

type of analysis (e.g., FF A determination, moi sture determination) in an analysis method;

additional customized analysis methods can be created using the method development

tool provided in the software. The analysis method specifies all relevant acquisition

parameters relating to the spectrometer and the auto-sampler, the component(s) to be

quantified, and the component method associated with each component, where the latter

specifies the spectral features that will be used for its quantification and the calibration

(linear or polynomial) by which the results will be computed. AlI results are archived

within the software and can be retrieved by referencing either the auto-sampler tray,

analysis method, operator that acquired the sampI es, value of result, etc. These results can

90

Page 108: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

be printed, exported to external software for further analysis or reporting, or re-processed

by a different analysis method. Through the use of eXtensible Markup Language (XML)

and the Microsoft.NET environment, UMPIRE Pro can relay information to a Laboratory

Information Management System (UMS).

Reagents. AlI solvents were of HPLC grade, obtained from Fisher Scientific (St.

Louis, MO). Acetonitrile was dried over 4-8 mesh 4A molecular sieves for a minimum of

two weeks and dispensed using a re-pipette (Hirschmann-Laborgerate, Germany)

protected by Drierite to prevent moisture ingress l6. The reagent solution employed in the

FFA analysis was prepared by dissolving NaHNCN in MeOH (2 g/L). This solution was

allowed to stand for ~4 days or until the v (C=N) band at 2100 cm- l completely

disappeared before use l5. Both extraction solvents (acetonitrile and methanol/NaHNCN)

were kept in sealed bottles and maintained over molecular sieves.

Calibration Standards. For the FF A calibration, calibration standards were

prepared by adding palmitic acid (0-1 %) to canola oil which had been passed through

activated silica gel column to remove any traces of FF A and other oxygenated

compounds. Moisture calibration standards were prepared by gravimetric addition of

distilled water (0-1000 ppm) to corn oil which had previously been kept over 10-20 mesh

Drierite (calcium sulfate) for at least one week.

Analytical Protocol. Both the FF A and moisture analyses are based on extraction

of the constituent of interest into an oil-immiscible solvent. For the analysis of FF A, the

extraction uses methanol containing a weak base, hydrogen cyanamide (NaHNCN), the

latter producing immediate conversion of acids into methanol-soluble salts without

causing saponification of the oils, whereas for moisture analysis the sample is extracted

with dry acetonitrile. The details associated with the procedures for these analyses have

been described elsewhereI5,16. Figure 6.1 presents a generalized schematic diagram of the

analytical proto col folIowed for both the calibration standards and the samples. For either

FF A or moi sture analyses, 20 g of oil was weighed into a tared 50-mL plastic vial to

within --±0.1 0 g. To the vial containing oil, 20 ml of the appropriate solvent was added

using a calibrated pro-pipette, and the vial was capped. The vials were agitated on a

vortex mixer for 30 s, and then loaded onto the autosampler tray. The thoroughly mixed

91

Page 109: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

/~~

samples were then left to stand for 10-15 min to ensure separation of the oil and solvent

layers. The 'automated FTIR spectral collection procedure was initiated with the

collection of an open-beam spectrum followed by a single-beam solvent spectrum, the

latter ratioed against the former to produce a reference solvent spectrum used to

determine cell path length. The single-beam spectrum of the solvent layer in each sample

vial was then recorded and ratioed against the single-beam solvent spectrum to obtain the

spectrum of the constituent being analyzed for; the same single-beam solvent spectrum

was used for a complete auto-sampler tray (56 samples). AlI spectra were collected by co­

adding 32 scans at a resolution of 8 cm-1 and a gain of 1.0.

20g Oil ..

20ml .. Solvent

! j ( Cap and ) Mix

t Spectrum 1 ~ [JlJL] (Background)

Let """ Separate

(----Â- ) i (Spectrum 2- Spectrum 1)

./ Upper

Spectrum 2 · (JuiL) -Layer (Sample)

Figure 6.1. General analytical protocol for FF A and H20 analysis of edible oils by FTIR spectroscopy.

For FFA analysis, the absorbance of the v (COO-) band at 1573 cm-1 relative to a

baseline of 1820 cm-1 was measured and the FF A content predicted from a calibration

equation obtained by regressing the corresponding absorbance values of the FF A

calibration standards against their FF A content, expressed as % FF A (oleic acid). For

moi sture analysis, the calibration equations used were based on relating the amount of

92

Page 110: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

moisture added to the moi sture standards to the absorbance of the strong v (O-H) band at

3629 cm-l and/or the weaker v (H-O-H) band at 1631 cm-l, both measured relative to a

single-point baseline at 2500 cm-l. We have established empirically that measurement of

the more sensitive 3629 cm-l band is optimal up to a moisture level of 2000 ppm, beyond

which the 1631 cm- l band is preferred. The 2000 ppm limit corresponds to an absorbance

at 3629 cm- l of ~0.7, and the switchover from the 3629 cm- l to the 1631 cm- l calibration

equation is automatically made by the software based on this absorbance criterion.

System Evaluation and Validation. Optimization of the analnical·system in terms

ofminimizing sample carryover was initially evaluated by sequentially analyzing low (L)

and high (H) FF A or moisture standards and calculating percent carryover as [(L2 -

L1)/H1] * 100. Sample carryover was further assessed by analyzing three replicates of six

different oils of unknown FF A or H20 content, these 18 samples being placed in the

autosampler tray in random order. Subsequently, reproducibility and accuracy of the

optimized automated FTIR method were assessed by standard addition of palmitic acid or

water to olive, sunflower, and peanut oils, each validation sample being analyzed in

duplicate, on different days, by the appropriate FTIR method and the results compared to

the gravimetrically added amounts of these constituents.

6.4. RESULTS AND DISCUSSION

6.4.1. System Overview and Software

Figure 6.2 illustrates the overall system configuration. Because the FF A and

moisture methods are designed to analyze low-viscosity solvent extracts rather than oils

per se, a small micro-pump can be used to aspirate the sample into the IR cell. Low dead

volume combined with 1/8" i.d. tubing minimizes sample volume (~15 ml required) and

allows the elimination of solvent rinses between sampI es. The sample vials have

conventional plastic septum caps but the silicone septum is replaced by a 20 mil Mylar

liner. This Mylar liner provides the hermetic seal required for moi sture analysis while

being readily punched through by the auto-sampler needle (Figure 6.3).

Both the FTIR spectrometer and the auto-sampler are controlled by UMPIRE Pro,

an upgrade of a proprietary software package designed as a general platform for

developing and implementing FTIR methods. Details regarding the design of the software

93

Page 111: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

E

D

Figure 6.2. The FTIRIauto-sampler system used for the analysis of FF A or moisture. The system is composed of micro pump (A), cell holder (B), FTIR spectrometer (C), autosampler tray (D) and robotic arm (E).

Figure 6.3. Auto-sampler needle punching through the Mylar liner.

94

Page 112: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

are presented in the Methodology section, and sorne features of the user interface are

illustrated in the panels of Figure 6.4. The user interface consists of a nested set of tab­

based menus that allow the user to implement calibration and method development

procedures. Figure 6.4A presents the "General" descriptive tab associated with the FF A

method, and Figure 6.4B presents the Scanning/Sampling tab, from which the FTIR

operating parameters and auto-sampler/pump parameters can be input. Figure 6.4C

illustrates the typical tabs included under the "Components" submenu, showing the tabs

that can be accessed from this submenu to define the peak height or area measurements to

Figure 6.4. The user interface of UMPlRE® Pro program. A and B, the "General" and "Scanning/Sampling" tabs for the "Modify Methods" submenu, respectively; C, "Peak! Area Position" tab for the "Components" submenu.

95

Page 113: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

be used for quantification and input calibration equations, waming messages for the

operator, altemate equations to be used under defined conditions, and baseline correction

procedures. This tab-based user interface makes it simple and intuitive to develop, assess,

and modify methods as well as calibration procedures. In addition to these development

and setup functions, spectra can be viewed, manipulated, and exported, and the analytical

result(s) can be displayed, printed, trended and/or associated with a client, machine or

operation as well as output to a Laboratory Information Management System (UMS) for

further processing or reporting.

For the present work, the autosampler was equipped with a 56-slot (7x8) tray. The

first two slots of the auto-sampler tray were reserved for the solvent used for the analysis

in question, the spectrum of which not only served as a background spectrum but also

was employed to determine the cell path length. For the latter purpose, a path length

calibration equation was developed by building three cells of different path lengths and

relating their path lengths, as determined from the fringe count in the spectrum of the

empty cell, to the height of an on-scale solvent band. Typical linear regression equations

obtained for acetonitrile and methanol, respectivelY' were:

Pathlength (/lm) = 673.51 * A 2627/2580 cm-1

- 18.488

R2 = 0.999 SD = 1.75 [6.1]

Pathlength (/lm) = 261.99* A 2044/1952 cm-1

- 0.251

R2 = 0.999 SD = 0.384 [6.2]

Using these equations, a cell path length is determined at the start of each run, of

which an ongoing record is maintained to monitor the condition of the cell over time.

This path length value is employed to normalize the spectral data subsequently collected

for the samples in the tray to a constant path length (500 and 100 /lm for H20 and FFA,

respectively). This procedure also allows one to rebuild cells without performing a

recalibration of the method.

96

Page 114: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

Once the auto-sampler tray is loaded with samples in capped vials, the whole tray

is shaken vigorously to extract the constituent of interest from the oil into the solvent and

then the samples are left to stand until phase separation occurs (~1 0 min). Automated

analysis of the samples is then initiated. For each successive sample in the tray, the

autosampler needle punches through the Mylar liner of the cap into the upper solvent

layer, which is then aspirated into the IR cell and its spectrum recorded. A "cell-full"

check is performed on the spectrum by measuring the height of a suitable solvent band in

a spectral region in which the component of interest does not absorb (generally, the same

band as used for the path length calculation); ifthis height is not within specifications, the

analysis is paused, and the operator is notified to check for air bubbles in the cell or

incomplete cellioading. When the "cell-full" check passes, the value of the parameter of

interest is automatically calculated from the calibration equation defined for the method.

Figure 6.5 and Figure 6.6 illustrate the differential spectra (solvent subtracted

out) obtained for moisture and FF A standards, respectively, after extraction with the

appropriate solvent. The calibration equations developed using these spectra were as

follows:

H20 (ppm) = 3418.70*A 3629cm-1 - 61.89

R2 = 0.998 SD=23.8 [6.3]

FFA (% w/w) = 5.05 * A 1573 cm-1 - 0.008

R2 = 0.999 SD = 0.0018 [6.4]

These calibration equations were obtained using calibration standards prepared by

spiking oils with the constituent of interest and extracting them with the appropriate

solvent in accordance with the analytical protocol employed for sample analysis.

However, since the spectral analysis is performed on the upper solvent layer containing

the extracted constituent, the matrix effects associated with the oils are effectively

eliminated, and hence the calibration procedure could be simplified substantially by

spiking the constituents directly into their respective solvents.

97

Page 115: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

A 0.3

0.2 ~ ~ = ~ .c ""' 0 <Il 0.1 .c ~

3600 i

2000

Wavenumber (cm-1)

C

D

i

1600

Figure 6.5. DifferentiaI spectra of canola oil standards spiked with water (0-2000 ppm) after extraction with acetonitrile. The major spectral features are the strong O-H stretching vibrations @ 3629 (A) and 3541 cm-1 (B), respectively, the ester carbonyl band due to traces oftriacylglycerol being extracted into acetonitrile @ 1740 cm-1 (C), and the weaker HOH bending vibration ofwater @ 1631 cm-1 (D).

A

0.2

0.00 -1----

1800 1700 1600 1500

Wavenumber (cm-1)

Figure 6.6. DifferentiaI spectra of canola oil standards spiked with palmitic acid (0.0-0.1 % w/w) after extraction with methanol containing NaHNCN. The spectra illustrate the response of the FFA salts produced which absorb @ 1573 cm-1 (A).

98

Page 116: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

'~',

6.4.2. System Evaluation and Validation

In optimizing the system parameters, the crucial variable is the pump time, the

objective being to minimize sample turnaround time white avoiding sample carryover.

The optimum pump time is largely a function of solvent viscosity and the wash volume

required to flush out the previous sample. Preliminary trials for setting the pump time

simply involved the analysis of high, intermediate and low FF A or moisture standards

placed in the auto-sampler tray in random order. Based on this screening work, it was

found that consistent results, without any carryover bias, were obtained using pump times

on the order of 30 s( corresponding to a sample volume of ~ 15 ml). with negligible

carryover «0.5%). Subsequent to these trials, a series of three replicate sets of six oils

(olive, sunflower, canola, sesame, grape seed, and safflower) of unknown FF A and

moisture content were prepared for analysis. These 18 samples were loaded into the auto­

sampler tray in random order for analysis by each method. Table 6.1 presents the

analytical data obtained for these runs. The excellent SD values confirm that sample

carryover was negligible.

Table 6.1. analyses)

FFA and Moisture Content of Various Edible Oils (mean of triplicate

OH

Canola

Extra virgin olive

Sunflower

Safflower

Sesame

Grape seed

%

0.019

0.270

0.024

0.024

1.170

0.025

FFA

sn 2.9*10-3

10* 10-3

2.9*10-3

5.8*10-3

2.6*10-3

10* 10-3

Moisture Content

ppm sn 39 1.9

850 12.3

145 9.8

87 5.2

302 1.9

168 12.0

Subsequently, a series of validation samples were prepared by gravimetrically

spiking refined olive, corn, peanut, and sunflower oils with random, but known amounts

ofwater (100-1000 ppm) and FFA (0.1-1%). These samples were analyzed in duplicate

using the appropriate FTIR method for each constituent. The FTIR data are plotted

against the amounts of moisture or FF A added to these oils in Figure 6.7 and Figure 6.8,

99

Page 117: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

<~,

respectively, and the corresponding zero-regression equations (forced through the origin)

are presented in Eqs. [6.5] and [6.6].

R2 = 0.997 SD = 14 ppm [6.5]

FTIR FFA = 1.011 FFA

R2 = 0.999 SD = 0.007 % FFA [6.6]

--"0 1.2 •• <J

* CI: ~ •• 1.0 QI -0

"t. 0.8 i ~ ~

.-0.6

~ .-

~ 0.4 "0 QI • == •• f e 0.2 :.. •• Qi .... QI

0.0 ; ~

0.0 0.2 0.4 0.6 0.8 1.0

Added FFA (w/w%, Oleic acid)

Figure 6.7. Plots for moi sture content vs. amounts added for the validation run. Error bars represent mean ± SD of three replicates.

100

Page 118: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

/..------.....

~,

700

~ 600 ~ e ! Q.

! Q. 500 '-"

0 i N 400

== !

"0 ! ~ 300 == .~

e 1- 200 1 ! ~ ; ~ 100

0 0 100 200 300 400 500 600

Added H20 (ppm)

Figure 6.8. Plots for FF A vs. amounts added for the validation run. Error bars represent mean ± SD of three replicates.

6.5. CONCLUSION

This study providesevidence that the automate d, high-speed analysis of edible

oils for FF A and moisture using a FTIR spectrometer coupled to an auto-sampler is

workable. Following sample preparation, which involves a simple in-vial extraction of

the constituent in question into a low-viscosity solvent, and loading of samples onto an

auto-sampler tray, the analysis is fully automated, with the FF A or moisture content

results being output on-screen or to a LIMS. When optimized, throughputs of ~ 100

samples/h can be achieved, which meets the requirements of high-volume contract,

process, or payment laboratories. It is also likely that the instrument configuration

described in this paper can be adapted to other FTIR-based methods such as the

determination of peroxide value or trans content.

101

Page 119: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

6.6. ACKNOWLEDGMENT

The authors would like to thank Sultan Qaboos University, Oman, and Thermal­

Lube Inc., Pointe-Claire, PQ, Canada for their financial support of this research. Thanks

also go to Dr. R Cocciardi for his technical assistance.

6.7. REFERENCES

1. O'Brien, RD., Fats and Oils: Formulating and Processing for Applications, Technomic Publishing Company, Lancaster, PA, 1998.

2. AOCS, Official Methods and Recommended Practices of the American Oil Chemists' Society, 4th edn., AOCS Press, Champaign, IL, 1998.

3. IUPAC, Standard Methodsfor the Analysis ofOils, Fats and Derivatives, 7th edn., Blackwell Scientific, London, 1992.

4. Ismail, A. A., van de Voort, F. R. and Sedman, J. Rapid Quantitative Determination of Free Fatty Acids in Fats and Oils by FTIR Spectroscopy. J. Am. Oil Chem. Soc., 70, 335-341 (1993).

5. AI-Alawi, A, van de Voort, F. R and Sedman, J. New Method for the Quantitative Determination of Free Fatty Acids in Oil by FTIR Spectroscopy. J. Am. Oil Chem. Soc., 81,441-446 (2004).

6. Li, H., van de Voort, F. R, Sedman, J. and Ismail, A. A. Rapid Determination of Cis and Trans Content, Iodine Value, and Saponification Number of Edible Oils by Fourier Transform Near-Infrared Spectroscopy. J. Am. Oil Chem. Soc., 76, 491-497 (1999).

7. van de Voort, F. R, Ismail, A. A. and Sedman, J. A Rapid, Automated Method for the Determination of Cis and Trans Content of Fats and Oils by Fourier Transform Infrared Spectroscopy. J. Am. Oil Chem. Soc., 72, 873-880 (1995).

8. Li, Hui; van de Voort, F. R, Ismail, A. A., Sedman, J. and Cox, R Trans Determination of Edible Oils by Fourier Transform Near-Infrared Spectroscopy. J. Am. Oil Chem. Soc., 77, 1061-1067 (2000).

9. Sedman, J., van De Voort, F. R and Ismail, A. A. Simultaneous Determination of Iodine Value and Trans Content of Fats and Oils by Single-Bounce Horizontal Attenuated Total Reflectance Fourier Transform Infrared Spectroscopy. J. Am. Oil Chem. Soc., 77, 399-403 (2000).

10. Dubois, J., van de Voort, F. R, Sedman, J., Ismail, A. A. and Ramaswamy, H. R Quantitative Fourier Transform Infrared Analysis for Anisidine Value and Aldehydes in Thermally Stressed Oils. J. Am. Oil Chem. Soc., 73, 787-794(1996).

11. van de Voort, F. R, Memon, K. P., Sedman, J. and Ismail, A. A. Determination of Solid Fat Index by Fourier-Transform Infrared Spectroscopy. J. Am. Oil Chem. Soc., 73, 411-416 (1996).

102

Page 120: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

~--/

12. van de Voort, F. R., Ismail, A. A., Sedman, J., Dubois, J and Nicodemo, T. The Determination of Peroxide Value by Fourier Transform Infrared (FTIR) Spectroscopy. J. Am. Di! Chem. Soc., 71 921-926 (1994).

13. Ma, K., Van De Voort, F. R., Sedman, J. and Ismail, A. A. Stoichiometric Determination of Hydroperoxides in Fats and Oils by Fourier Transform Infrared Spectroscopy. J. Am. Di! Chem. Soc., 74, 897-906 (1997).

14. Li, H., Van de Voort, F. R., Ismail, A. A. and Cox, R. Determination ofPeroxide Value by Fourier Transform Near-Infrared Spectroscopy. J. Am. Di! Chem. Soc., 77, 137-142 (2000).

15. Al-Alawi, A., van de Voort, F. R. and Sedman, J. A New FTIR Method for the Analysis ofLow Levels ofFFA in Refined Edible Oils. Spectrosc. Lett. 38,389-403 (2005).

16. Al-Alawi, A., van de Voort, F. R. and Sedman, J. A New FTIR Method for the Determination of Low Levels of Moisture in Edible Oils. Appl. Spectrosc., (in press, 2005).

103

Page 121: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

CHAPTER 7

GENERAL CONCLUSION

The methods documented in this thesis add a significant portion to the continuing

work of FTIR methodology development in oil analysis. Each part of the work laid out in

this thesis provides a solution to a cornrnon problem in its field or presents an improved

me ans of carrying out edible oil analysis. The first method described in Chapter 3

provides a novel and yet workable approach to overcome matrix effects which for years

hindered development of a general method for free fatty acids analysis of edible oils. The

method uses a suspension of a weak base, potassium phthalimide (K-phthal) in 1-

propanol (l-PrOH), to convert FF As present in oils to their carboxylate salts without

causing oil saponification. The matrix effects were resolved by splitting the diluted oil

samples into two halves, with one half treated with the K-phthal reagent and the other

halfwith I-PrOH (blank reagent). The spectra of the two halves were then coUected, and

a differential spectrurn obtained to ratio out the invariant spectral contributions from the

oil sarnple. Quantification of FF A was based on direct measurement of the peak height of

the free fatty acids salt band at 1570 cm-1. Although this new method did not provide a

substantial gain in terms of sensitivity (~ 0.2% FF A) over the standard and other

titrimetric methods, it offers both specificity to free fatty acids and analysis independent

of oil type, with differential spectroscopy being employed to circumvent matrix effects.

Moreover, unlike the titrimetric methods, thi~ FTIR method does not require reagent

standardization, electrode conditioning or maintenance.

The subsequent FF A method described in Chapter 4, takes a different approach to

the sarne analysis that makes it simpler to perforrn, more rapid and more sensitive without

losing specificity. The method is extraction-based and the target constituent (FF A in the

forrn of its salt) is extracted from oil sarnples using a polar oil-immiscible solvent

(methanol) containing a weak base (sodium hydrogen cyanamide) which is soluble in the

extracting solvent. Not only does this simplify the extraction procedure but it also

significantly enhances sensitivity and obviates oil matrix effects as the oil is no longer

present in the sarnple being analyzed spectroscopically. This method provides a simple

and robust means by which to analyze for low levels of FF A with a high degree of

104

Page 122: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

accuracy (± 0.0006%) with low solvent consumption (2 ml/analysis). Furthermore, with a

simple adjustment of the reagent:oil ratio, the method is scalable for the analysis of semi­

refined and crude oils containing up to 4.0% FFA. An additional benefit ofthis method is

that its sensitivity opens the door to more accurate monitoring of the refining processes

carried out on edible oils and the possible use of FF As as an indicator of lipid oxidation.

In Chapter 5, the sarne principle of extraction and concentration was used to

develop a rapid, practical, and accurate FTIR method for the determination of moisture

content in edible oils. This was possible because of the high polarity of water, its strong

mid IR absorption as weIl as the unique solvating characteristics of acetonitrile obviating

the need to carry out any conversion reactions to pro duce more IR-active species. For the

greatest sensitivity (0-200 ppm) the -OH stretching band (3629 cm- I) is used, while for

higher moisture levels (200-2000 ppm) the H-O-H bending band (1631 cm- I) of water is

suitable. One unique feature of the method is that the acetonitrile solvent does not need to

be absolutely dry to obtain accurate results, as the method corrects for traces of moi sture

present via spectral ratioing against the acetonitrile used to extract the sarnple, greatly

simplifying the methodology. It was also shown that a1cohols, hydroperoxides, and FF As,

which may be extracted into acetonitrile, do not interfere with the analysis. This robust

method has the potential to be a practical alternative to the Karl Fischer method, which

has many variables associated with its analytical performance.

The work in Chapter 6 concerns the implementation of the FF A and moi sture

extraction-based methods on an FTIR spectrometer coupled to an auto-sarnpler to

automate these analyses. The system developed effectively automates sample loading,

spectral collection and calculation of results using pre-prograrnmed algorithrns, with

output to screen or to a Laboratory Information Management System. It was tested,

optimized and validated, and shown to be capable of sarnple throughputs of >60

sarnpleslh with an accuracy very similar to that attained by using the manual procedures.

This system is designed for high-volurne contract, process or payment laboratories and

lays the groundwork for extending its utility to other common edible oil analyses, such as

peroxide value or trans content.

105

Page 123: Ahmed Ali AI-Alawi September 2005 · ©Ahmed Ali AI-Alawi, 2005 . 1+1 Library and Archives Canada Bibliothèque et Archives Canada Published Heritage Branch Direction du Patrimoine

The development of the methods described in this thesis has been based on the

design of novel approaches to problems that have hindered the full exploitation of the

power of FTIR spectroscopy as a rapid and simple instrumental tool for edible oil

analysis. This research has provided very accurate and specific measurements for two of

the most important quality factors associated with edible oils. FF A and moisture methods

have been developed that are superior in specificity and simplicity to their respective

conventional titrimetric methods while providing comparable or better sensitivity and

accuracy. Furthermore, the integration of an FTIR spectrometer with an auto-sampler,

pump and appropriate software has made automated analysis of edible oils for FF A and

moi sture a realistic possibility. The adoption of such methods by industry can have a

significant impact on the monitoring of oil processing operations and product quality

control.

106